This is the Difference Between a Hypothesis and a Theory

What to Know A hypothesis is an assumption made before any research has been done. It is formed so that it can be tested to see if it might be true. A theory is a principle formed to explain the things already shown in data. Because of the rigors of experiment and control, it is much more likely that a theory will be true than a hypothesis.

As anyone who has worked in a laboratory or out in the field can tell you, science is about process: that of observing, making inferences about those observations, and then performing tests to see if the truth value of those inferences holds up. The scientific method is designed to be a rigorous procedure for acquiring knowledge about the world around us.

hypothesis

In scientific reasoning, a hypothesis is constructed before any applicable research has been done. A theory, on the other hand, is supported by evidence: it's a principle formed as an attempt to explain things that have already been substantiated by data.

Toward that end, science employs a particular vocabulary for describing how ideas are proposed, tested, and supported or disproven. And that's where we see the difference between a hypothesis and a theory .

A hypothesis is an assumption, something proposed for the sake of argument so that it can be tested to see if it might be true.

In the scientific method, the hypothesis is constructed before any applicable research has been done, apart from a basic background review. You ask a question, read up on what has been studied before, and then form a hypothesis.

What is a Hypothesis?

A hypothesis is usually tentative, an assumption or suggestion made strictly for the objective of being tested.

When a character which has been lost in a breed, reappears after a great number of generations, the most probable hypothesis is, not that the offspring suddenly takes after an ancestor some hundred generations distant, but that in each successive generation there has been a tendency to reproduce the character in question, which at last, under unknown favourable conditions, gains an ascendancy. Charles Darwin, On the Origin of Species , 1859 According to one widely reported hypothesis , cell-phone transmissions were disrupting the bees' navigational abilities. (Few experts took the cell-phone conjecture seriously; as one scientist said to me, "If that were the case, Dave Hackenberg's hives would have been dead a long time ago.") Elizabeth Kolbert, The New Yorker , 6 Aug. 2007

What is a Theory?

A theory , in contrast, is a principle that has been formed as an attempt to explain things that have already been substantiated by data. It is used in the names of a number of principles accepted in the scientific community, such as the Big Bang Theory . Because of the rigors of experimentation and control, its likelihood as truth is much higher than that of a hypothesis.

It is evident, on our theory , that coasts merely fringed by reefs cannot have subsided to any perceptible amount; and therefore they must, since the growth of their corals, either have remained stationary or have been upheaved. Now, it is remarkable how generally it can be shown, by the presence of upraised organic remains, that the fringed islands have been elevated: and so far, this is indirect evidence in favour of our theory . Charles Darwin, The Voyage of the Beagle , 1839 An example of a fundamental principle in physics, first proposed by Galileo in 1632 and extended by Einstein in 1905, is the following: All observers traveling at constant velocity relative to one another, should witness identical laws of nature. From this principle, Einstein derived his theory of special relativity. Alan Lightman, Harper's , December 2011

Non-Scientific Use

In non-scientific use, however, hypothesis and theory are often used interchangeably to mean simply an idea, speculation, or hunch (though theory is more common in this regard):

The theory of the teacher with all these immigrant kids was that if you spoke English loudly enough they would eventually understand. E. L. Doctorow, Loon Lake , 1979 Chicago is famous for asking questions for which there can be no boilerplate answers. Example: given the probability that the federal tax code, nondairy creamer, Dennis Rodman and the art of mime all came from outer space, name something else that has extraterrestrial origins and defend your hypothesis . John McCormick, Newsweek , 5 Apr. 1999 In his mind's eye, Miller saw his case suddenly taking form: Richard Bailey had Helen Brach killed because she was threatening to sue him over the horses she had purchased. It was, he realized, only a theory , but it was one he felt certain he could, in time, prove. Full of urgency, a man with a mission now that he had a hypothesis to guide him, he issued new orders to his troops: Find out everything you can about Richard Bailey and his crowd. Howard Blum, Vanity Fair , January 1995

And sometimes one term is used as a genus, or a means for defining the other:

Laplace's popular version of his astronomy, the Système du monde , was famous for introducing what came to be known as the nebular hypothesis , the theory that the solar system was formed by the condensation, through gradual cooling, of the gaseous atmosphere (the nebulae) surrounding the sun. Louis Menand, The Metaphysical Club , 2001 Researchers use this information to support the gateway drug theory — the hypothesis that using one intoxicating substance leads to future use of another. Jordy Byrd, The Pacific Northwest Inlander , 6 May 2015 Fox, the business and economics columnist for Time magazine, tells the story of the professors who enabled those abuses under the banner of the financial theory known as the efficient market hypothesis . Paul Krugman, The New York Times Book Review , 9 Aug. 2009

Incorrect Interpretations of "Theory"

Since this casual use does away with the distinctions upheld by the scientific community, hypothesis and theory are prone to being wrongly interpreted even when they are encountered in scientific contexts—or at least, contexts that allude to scientific study without making the critical distinction that scientists employ when weighing hypotheses and theories.

The most common occurrence is when theory is interpreted—and sometimes even gleefully seized upon—to mean something having less truth value than other scientific principles. (The word law applies to principles so firmly established that they are almost never questioned, such as the law of gravity.)

This mistake is one of projection: since we use theory in general use to mean something lightly speculated, then it's implied that scientists must be talking about the same level of uncertainty when they use theory to refer to their well-tested and reasoned principles.

The distinction has come to the forefront particularly on occasions when the content of science curricula in schools has been challenged—notably, when a school board in Georgia put stickers on textbooks stating that evolution was "a theory, not a fact, regarding the origin of living things." As Kenneth R. Miller, a cell biologist at Brown University, has said , a theory "doesn’t mean a hunch or a guess. A theory is a system of explanations that ties together a whole bunch of facts. It not only explains those facts, but predicts what you ought to find from other observations and experiments.”

While theories are never completely infallible, they form the basis of scientific reasoning because, as Miller said "to the best of our ability, we’ve tested them, and they’ve held up."

More Differences Explained

  • Epidemic vs. Pandemic
  • Diagnosis vs. Prognosis
  • Treatment vs. Cure

Word of the Day

See Definitions and Examples »

Get Word of the Day daily email!

Games & Quizzes

Play Quordle: Guess all four words in a limited number of tries.  Each of your guesses must be a real 5-letter word.

Commonly Confused

'canceled' or 'cancelled', 'virus' vs. 'bacteria', your vs. you're: how to use them correctly, is it 'jail' or 'prison', 'deduction' vs. 'induction' vs. 'abduction', grammar & usage, 7 pairs of commonly confused words, did we change the definition of 'literally', more commonly mispronounced words, the tangled history of 'it's' and 'its', more commonly misspelled words, 10 bird names that sound like insults (and sometimes are), eavesdrop, fiasco, and 8 more words with surprising origins, 'when pigs fly' and other barnyard idioms, the words of the week - apr. 5, 12 bird names that sound like compliments.

Back Home

  • Science Notes Posts
  • Contact Science Notes
  • Todd Helmenstine Biography
  • Anne Helmenstine Biography
  • Free Printable Periodic Tables (PDF and PNG)
  • Periodic Table Wallpapers
  • Interactive Periodic Table
  • Periodic Table Posters
  • How to Grow Crystals
  • Chemistry Projects
  • Fire and Flames Projects
  • Holiday Science
  • Chemistry Problems With Answers
  • Physics Problems
  • Unit Conversion Example Problems
  • Chemistry Worksheets
  • Biology Worksheets
  • Periodic Table Worksheets
  • Physical Science Worksheets
  • Science Lab Worksheets
  • My Amazon Books

Scientific Theory Definition and Examples

Scientific Theory Definition

A scientific theory is a well-established explanation of some aspect of the natural world. Theories come from scientific data and multiple experiments. While it is not possible to prove a theory, a single contrary result using the scientific method can disprove it. In other words, a theory is testable and falsifiable.

Examples of Scientific Theories

There are many scientific theory in different disciplines:

  • Astronomy : theory of stellar nucleosynthesis , theory of stellar evolution
  • Biology : cell theory, theory of evolution, germ theory, dual inheritance theory
  • Chemistry : atomic theory, Bronsted Lowry acid-base theory , kinetic molecular theory of gases , Lewis acid-base theory , molecular theory, valence bond theory
  • Geology : climate change theory, plate tectonics theory
  • Physics : Big Bang theory, perturbation theory, theory of relativity, quantum field theory

Criteria for a Theory

In order for an explanation of the natural world to be a theory, it meets certain criteria:

  • A theory is falsifiable. At some point, a theory withstands testing and experimentation using the scientific method.
  • A theory is supported by lots of independent evidence.
  • A theory explains existing experimental results and predicts outcomes of new experiments at least as well as other theories.

Difference Between a Scientific Theory and Theory

Usually, a scientific theory is just called a theory. However, a theory in science means something different from the way most people use the word. For example, if frogs rain down from the sky, a person might observe the frogs and say, “I have a theory about why that happened.” While that theory might be an explanation, it is not based on multiple observations and experiments. It might not be testable and falsifiable. It’s not a scientific theory (although it could eventually become one).

Value of Disproven Theories

Even though some theories are incorrect, they often retain value.

For example, Arrhenius acid-base theory does not explain the behavior of chemicals lacking hydrogen that behave as acids. The Bronsted Lowry and Lewis theories do a better job of explaining this behavior. Yet, the Arrhenius theory predicts the behavior of most acids and is easier for people to understand.

Another example is the theory of Newtonian mechanics. The theory of relativity is much more inclusive than Newtonian mechanics, which breaks down in certain frames of reference or at speeds close to the speed of light . But, Newtonian mechanics is much simpler to understand and its equations apply to everyday behavior.

Difference Between a Scientific Theory and a Scientific Law

The scientific method leads to the formulation of both scientific theories and laws . Both theories and laws are falsifiable. Both theories and laws help with making predictions about the natural world. However, there is a key difference.

A theory explains why or how something works, while a law describes what happens without explaining it. Often, you see laws written in the form of equations or formulas.

Theories and laws are related, but theories never become laws or vice versa.

Theory vs Hypothesis

A hypothesis is a proposition that is tested via an experiment. A theory results from many, many tested hypotheses.

Theory vs Fact

Theories depend on facts, but the two words mean different things. A fact is an irrefutable piece of evidence or data. Facts never change. A theory, on the other hand, may be modified or disproven.

Difference Between a Theory and a Model

Both theories and models allow a scientist to form a hypothesis and make predictions about future outcomes. However, a theory both describes and explains, while a model only describes. For example, a model of the solar system shows the arrangement of planets and asteroids in a plane around the Sun, but it does not explain how or why they got into their positions.

  • Frigg, Roman (2006). “ Scientific Representation and the Semantic View of Theories .”  Theoria . 55 (2): 183–206. 
  • Halvorson, Hans (2012). “What Scientific Theories Could Not Be.”  Philosophy of Science . 79 (2): 183–206. doi: 10.1086/664745
  • McComas, William F. (December 30, 2013).  The Language of Science Education: An Expanded Glossary of Key Terms and Concepts in Science Teaching and Learning . Springer Science & Business Media. ISBN 978-94-6209-497-0.
  • National Academy of Sciences (US) (1999). Science and Creationism: A View from the National Academy of Sciences (2nd ed.). National Academies Press. doi: 10.17226/6024  ISBN 978-0-309-06406-4. 
  • Suppe, Frederick (1998). “Understanding Scientific Theories: An Assessment of Developments, 1969–1998.”  Philosophy of Science . 67: S102–S115. doi: 10.1086/392812

Related Posts

Hypothesis vs. Theory

A hypothesis is either a suggested explanation for an observable phenomenon, or a reasoned prediction of a possible causal correlation among multiple phenomena. In science , a theory is a tested, well-substantiated, unifying explanation for a set of verified, proven factors. A theory is always backed by evidence; a hypothesis is only a suggested possible outcome, and is testable and falsifiable.

Comparison chart

Examples of theory and hypothesis.

Theory: Einstein's theory of relativity is a theory because it has been tested and verified innumerable times, with results consistently verifying Einstein's conclusion. However, simply because Einstein's conclusion has become a theory does not mean testing of this theory has stopped; all science is ongoing. See also the Big Bang theory , germ theory , and climate change .

Hypothesis: One might think that a prisoner who learns a work skill while in prison will be less likely to commit a crime when released. This is a hypothesis, an "educated guess." The scientific method can be used to test this hypothesis, to either prove it is false or prove that it warrants further study. (Note: Simply because a hypothesis is not found to be false does not mean it is true all or even most of the time. If it is consistently true after considerable time and research, it may be on its way to becoming a theory.)

This video further explains the difference between a theory and a hypothesis:

Common Misconception

People often tend to say "theory" when what they're actually talking about is a hypothesis. For instance, "Migraines are caused by drinking coffee after 2 p.m. — well, it's just a theory, not a rule."

This is actually a logically reasoned proposal based on an observation — say 2 instances of drinking coffee after 2 p.m. caused a migraine — but even if this were true, the migraine could have actually been caused by some other factors.

Because this observation is merely a reasoned possibility, it is testable and can be falsified — which makes it a hypothesis, not a theory.

  • What is a Scientific Hypothesis? - LiveScience
  • Wikipedia:Scientific theory

Related Comparisons

Accuracy vs Precision

Share this comparison via:

If you read this far, you should follow us:

"Hypothesis vs Theory." Diffen.com. Diffen LLC, n.d. Web. 3 Mar 2024. < >

Comments: Hypothesis vs Theory

Anonymous comments (2).

October 11, 2013, 1:11pm "In science, a theory is a well-substantiated, unifying explanation for a set of verified, proven hypotheses." But there's no such thing as "proven hypotheses". Hypotheses can be tested/falsified, they can't be "proven". That's just not how science works. Logical deductions based on axioms can be proven, but not scientific hypotheses. On top of that I find it somewhat strange to claim that a theory doesn't have to be testable, if it's built up from hypotheses, which DO have to be testable... — 80.✗.✗.139
May 6, 2014, 11:45pm "Evolution is a theory, not a fact, regarding the origin of living things." this statement is poorly formed because it implies that a thing is a theory until it gets proven and then it is somehow promoted to fact. this is just a misunderstanding of what the words mean, and of how science progresses generally. to say that a theory is inherently dubious because "it isn't a fact" is pretty much a meaningless statement. no expression which qualified as a mere fact could do a very good job of explaining the complicated process by which species have arisen on Earth over the last billion years. in fact, if you claimed that you could come up with such a single fact, now THAT would be dubious! everything we observe in nature supports the theory of evolution, and nothing we observe contradicts it. when you can say this about a theory, it's a pretty fair bet that the theory is correct. — 71.✗.✗.151
  • Accuracy vs Precision
  • Deductive vs Inductive
  • Subjective vs Objective
  • Subconscious vs Unconscious mind
  • Qualitative vs Quantitative
  • Creationism vs Evolution

Edit or create new comparisons in your area of expertise.

Stay connected

© All rights reserved.

If you're seeing this message, it means we're having trouble loading external resources on our website.

If you're behind a web filter, please make sure that the domains *.kastatic.org and *.kasandbox.org are unblocked.

To log in and use all the features of Khan Academy, please enable JavaScript in your browser.

Biology library

Course: biology library   >   unit 1, the scientific method.

  • Controlled experiments
  • The scientific method and experimental design

Introduction

  • Make an observation.
  • Ask a question.
  • Form a hypothesis , or testable explanation.
  • Make a prediction based on the hypothesis.
  • Test the prediction.
  • Iterate: use the results to make new hypotheses or predictions.

Scientific method example: Failure to toast

1. make an observation..

  • Observation: the toaster won't toast.

2. Ask a question.

  • Question: Why won't my toaster toast?

3. Propose a hypothesis.

  • Hypothesis: Maybe the outlet is broken.

4. Make predictions.

  • Prediction: If I plug the toaster into a different outlet, then it will toast the bread.

5. Test the predictions.

  • Test of prediction: Plug the toaster into a different outlet and try again.
  • If the toaster does toast, then the hypothesis is supported—likely correct.
  • If the toaster doesn't toast, then the hypothesis is not supported—likely wrong.

Logical possibility

Practical possibility, building a body of evidence, 6. iterate..

  • Iteration time!
  • If the hypothesis was supported, we might do additional tests to confirm it, or revise it to be more specific. For instance, we might investigate why the outlet is broken.
  • If the hypothesis was not supported, we would come up with a new hypothesis. For instance, the next hypothesis might be that there's a broken wire in the toaster.

Want to join the conversation?

  • Upvote Button navigates to signup page
  • Downvote Button navigates to signup page
  • Flag Button navigates to signup page

Incredible Answer

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
  • J Korean Med Sci
  • v.34(45); 2019 Nov 25

Logo of jkms

Scientific Hypotheses: Writing, Promoting, and Predicting Implications

Armen yuri gasparyan.

1 Departments of Rheumatology and Research and Development, Dudley Group NHS Foundation Trust (Teaching Trust of the University of Birmingham, UK), Russells Hall Hospital, Dudley, West Midlands, UK.

Lilit Ayvazyan

2 Department of Medical Chemistry, Yerevan State Medical University, Yerevan, Armenia.

Ulzhan Mukanova

3 Department of Surgical Disciplines, South Kazakhstan Medical Academy, Shymkent, Kazakhstan.

Marlen Yessirkepov

4 Department of Biology and Biochemistry, South Kazakhstan Medical Academy, Shymkent, Kazakhstan.

George D. Kitas

5 Arthritis Research UK Epidemiology Unit, University of Manchester, Manchester, UK.

Scientific hypotheses are essential for progress in rapidly developing academic disciplines. Proposing new ideas and hypotheses require thorough analyses of evidence-based data and predictions of the implications. One of the main concerns relates to the ethical implications of the generated hypotheses. The authors may need to outline potential benefits and limitations of their suggestions and target widely visible publication outlets to ignite discussion by experts and start testing the hypotheses. Not many publication outlets are currently welcoming hypotheses and unconventional ideas that may open gates to criticism and conservative remarks. A few scholarly journals guide the authors on how to structure hypotheses. Reflecting on general and specific issues around the subject matter is often recommended for drafting a well-structured hypothesis article. An analysis of influential hypotheses, presented in this article, particularly Strachan's hygiene hypothesis with global implications in the field of immunology and allergy, points to the need for properly interpreting and testing new suggestions. Envisaging the ethical implications of the hypotheses should be considered both by authors and journal editors during the writing and publishing process.

INTRODUCTION

We live in times of digitization that radically changes scientific research, reporting, and publishing strategies. Researchers all over the world are overwhelmed with processing large volumes of information and searching through numerous online platforms, all of which make the whole process of scholarly analysis and synthesis complex and sophisticated.

Current research activities are diversifying to combine scientific observations with analysis of facts recorded by scholars from various professional backgrounds. 1 Citation analyses and networking on social media are also becoming essential for shaping research and publishing strategies globally. 2 Learning specifics of increasingly interdisciplinary research studies and acquiring information facilitation skills aid researchers in formulating innovative ideas and predicting developments in interrelated scientific fields.

Arguably, researchers are currently offered more opportunities than in the past for generating new ideas by performing their routine laboratory activities, observing individual cases and unusual developments, and critically analyzing published scientific facts. What they need at the start of their research is to formulate a scientific hypothesis that revisits conventional theories, real-world processes, and related evidence to propose new studies and test ideas in an ethical way. 3 Such a hypothesis can be of most benefit if published in an ethical journal with wide visibility and exposure to relevant online databases and promotion platforms.

Although hypotheses are crucially important for the scientific progress, only few highly skilled researchers formulate and eventually publish their innovative ideas per se . Understandably, in an increasingly competitive research environment, most authors would prefer to prioritize their ideas by discussing and conducting tests in their own laboratories or clinical departments, and publishing research reports afterwards. However, there are instances when simple observations and research studies in a single center are not capable of explaining and testing new groundbreaking ideas. Formulating hypothesis articles first and calling for multicenter and interdisciplinary research can be a solution in such instances, potentially launching influential scientific directions, if not academic disciplines.

The aim of this article is to overview the importance and implications of infrequently published scientific hypotheses that may open new avenues of thinking and research.

Despite the seemingly established views on innovative ideas and hypotheses as essential research tools, no structured definition exists to tag the term and systematically track related articles. In 1973, the Medical Subject Heading (MeSH) of the U.S. National Library of Medicine introduced “Research Design” as a structured keyword that referred to the importance of collecting data and properly testing hypotheses, and indirectly linked the term to ethics, methods and standards, among many other subheadings.

One of the experts in the field defines “hypothesis” as a well-argued analysis of available evidence to provide a realistic (scientific) explanation of existing facts, fill gaps in public understanding of sophisticated processes, and propose a new theory or a test. 4 A hypothesis can be proven wrong partially or entirely. However, even such an erroneous hypothesis may influence progress in science by initiating professional debates that help generate more realistic ideas. The main ethical requirement for hypothesis authors is to be honest about the limitations of their suggestions. 5

EXAMPLES OF INFLUENTIAL SCIENTIFIC HYPOTHESES

Daily routine in a research laboratory may lead to groundbreaking discoveries provided the daily accounts are comprehensively analyzed and reproduced by peers. The discovery of penicillin by Sir Alexander Fleming (1928) can be viewed as a prime example of such discoveries that introduced therapies to treat staphylococcal and streptococcal infections and modulate blood coagulation. 6 , 7 Penicillin got worldwide recognition due to the inventor's seminal works published by highly prestigious and widely visible British journals, effective ‘real-world’ antibiotic therapy of pneumonia and wounds during World War II, and euphoric media coverage. 8 In 1945, Fleming, Florey and Chain got a much deserved Nobel Prize in Physiology or Medicine for the discovery that led to the mass production of the wonder drug in the U.S. and ‘real-world practice’ that tested the use of penicillin. What remained globally unnoticed is that Zinaida Yermolyeva, the outstanding Soviet microbiologist, created the Soviet penicillin, which turned out to be more effective than the Anglo-American penicillin and entered mass production in 1943; that year marked the turning of the tide of the Great Patriotic War. 9 One of the reasons of the widely unnoticed discovery of Zinaida Yermolyeva is that her works were published exclusively by local Russian (Soviet) journals.

The past decades have been marked by an unprecedented growth of multicenter and global research studies involving hundreds and thousands of human subjects. This trend is shaped by an increasing number of reports on clinical trials and large cohort studies that create a strong evidence base for practice recommendations. Mega-studies may help generate and test large-scale hypotheses aiming to solve health issues globally. Properly designed epidemiological studies, for example, may introduce clarity to the hygiene hypothesis that was originally proposed by David Strachan in 1989. 10 David Strachan studied the epidemiology of hay fever in a cohort of 17,414 British children and concluded that declining family size and improved personal hygiene had reduced the chances of cross infections in families, resulting in epidemics of atopic disease in post-industrial Britain. Over the past four decades, several related hypotheses have been proposed to expand the potential role of symbiotic microorganisms and parasites in the development of human physiological immune responses early in life and protection from allergic and autoimmune diseases later on. 11 , 12 Given the popularity and the scientific importance of the hygiene hypothesis, it was introduced as a MeSH term in 2012. 13

Hypotheses can be proposed based on an analysis of recorded historic events that resulted in mass migrations and spreading of certain genetic diseases. As a prime example, familial Mediterranean fever (FMF), the prototype periodic fever syndrome, is believed to spread from Mesopotamia to the Mediterranean region and all over Europe due to migrations and religious prosecutions millennia ago. 14 Genetic mutations spearing mild clinical forms of FMF are hypothesized to emerge and persist in the Mediterranean region as protective factors against more serious infectious diseases, particularly tuberculosis, historically common in that part of the world. 15 The speculations over the advantages of carrying the MEditerranean FeVer (MEFV) gene are further strengthened by recorded low mortality rates from tuberculosis among FMF patients of different nationalities living in Tunisia in the first half of the 20th century. 16

Diagnostic hypotheses shedding light on peculiarities of diseases throughout the history of mankind can be formulated using artefacts, particularly historic paintings. 17 Such paintings may reveal joint deformities and disfigurements due to rheumatic diseases in individual subjects. A series of paintings with similar signs of pathological conditions interpreted in a historic context may uncover mysteries of epidemics of certain diseases, which is the case with Ruben's paintings depicting signs of rheumatic hands and making some doctors to believe that rheumatoid arthritis was common in Europe in the 16th and 17th century. 18

WRITING SCIENTIFIC HYPOTHESES

There are author instructions of a few journals that specifically guide how to structure, format, and make submissions categorized as hypotheses attractive. One of the examples is presented by Med Hypotheses , the flagship journal in its field with more than four decades of publishing and influencing hypothesis authors globally. However, such guidance is not based on widely discussed, implemented, and approved reporting standards, which are becoming mandatory for all scholarly journals.

Generating new ideas and scientific hypotheses is a sophisticated task since not all researchers and authors are skilled to plan, conduct, and interpret various research studies. Some experience with formulating focused research questions and strong working hypotheses of original research studies is definitely helpful for advancing critical appraisal skills. However, aspiring authors of scientific hypotheses may need something different, which is more related to discerning scientific facts, pooling homogenous data from primary research works, and synthesizing new information in a systematic way by analyzing similar sets of articles. To some extent, this activity is reminiscent of writing narrative and systematic reviews. As in the case of reviews, scientific hypotheses need to be formulated on the basis of comprehensive search strategies to retrieve all available studies on the topics of interest and then synthesize new information selectively referring to the most relevant items. One of the main differences between scientific hypothesis and review articles relates to the volume of supportive literature sources ( Table 1 ). In fact, hypothesis is usually formulated by referring to a few scientific facts or compelling evidence derived from a handful of literature sources. 19 By contrast, reviews require analyses of a large number of published documents retrieved from several well-organized and evidence-based databases in accordance with predefined search strategies. 20 , 21 , 22

The format of hypotheses, especially the implications part, may vary widely across disciplines. Clinicians may limit their suggestions to the clinical manifestations of diseases, outcomes, and management strategies. Basic and laboratory scientists analysing genetic, molecular, and biochemical mechanisms may need to view beyond the frames of their narrow fields and predict social and population-based implications of the proposed ideas. 23

Advanced writing skills are essential for presenting an interesting theoretical article which appeals to the global readership. Merely listing opposing facts and ideas, without proper interpretation and analysis, may distract the experienced readers. The essence of a great hypothesis is a story behind the scientific facts and evidence-based data.

ETHICAL IMPLICATIONS

The authors of hypotheses substantiate their arguments by referring to and discerning rational points from published articles that might be overlooked by others. Their arguments may contradict the established theories and practices, and pose global ethical issues, particularly when more or less efficient medical technologies and public health interventions are devalued. The ethical issues may arise primarily because of the careless references to articles with low priorities, inadequate and apparently unethical methodologies, and concealed reporting of negative results. 24 , 25

Misinterpretation and misunderstanding of the published ideas and scientific hypotheses may complicate the issue further. For example, Alexander Fleming, whose innovative ideas of penicillin use to kill susceptible bacteria saved millions of lives, warned of the consequences of uncontrolled prescription of the drug. The issue of antibiotic resistance had emerged within the first ten years of penicillin use on a global scale due to the overprescription that affected the efficacy of antibiotic therapies, with undesirable consequences for millions. 26

The misunderstanding of the hygiene hypothesis that primarily aimed to shed light on the role of the microbiome in allergic and autoimmune diseases resulted in decline of public confidence in hygiene with dire societal implications, forcing some experts to abandon the original idea. 27 , 28 Although that hypothesis is unrelated to the issue of vaccinations, the public misunderstanding has resulted in decline of vaccinations at a time of upsurge of old and new infections.

A number of ethical issues are posed by the denial of the viral (human immunodeficiency viruses; HIV) hypothesis of acquired Immune deficiency Syndrome (AIDS) by Peter Duesberg, who overviewed the links between illicit recreational drugs and antiretroviral therapies with AIDS and refuted the etiological role of HIV. 29 That controversial hypothesis was rejected by several journals, but was eventually published without external peer review at Med Hypotheses in 2010. The publication itself raised concerns of the unconventional editorial policy of the journal, causing major perturbations and more scrutinized publishing policies by journals processing hypotheses.

WHERE TO PUBLISH HYPOTHESES

Although scientific authors are currently well informed and equipped with search tools to draft evidence-based hypotheses, there are still limited quality publication outlets calling for related articles. The journal editors may be hesitant to publish articles that do not adhere to any research reporting guidelines and open gates for harsh criticism of unconventional and untested ideas. Occasionally, the editors opting for open-access publishing and upgrading their ethics regulations launch a section to selectively publish scientific hypotheses attractive to the experienced readers. 30 However, the absence of approved standards for this article type, particularly no mandate for outlining potential ethical implications, may lead to publication of potentially harmful ideas in an attractive format.

A suggestion of simultaneously publishing multiple or alternative hypotheses to balance the reader views and feedback is a potential solution for the mainstream scholarly journals. 31 However, that option alone is hardly applicable to emerging journals with unconventional quality checks and peer review, accumulating papers with multiple rejections by established journals.

A large group of experts view hypotheses with improbable and controversial ideas publishable after formal editorial (in-house) checks to preserve the authors' genuine ideas and avoid conservative amendments imposed by external peer reviewers. 32 That approach may be acceptable for established publishers with large teams of experienced editors. However, the same approach can lead to dire consequences if employed by nonselective start-up, open-access journals processing all types of articles and primarily accepting those with charged publication fees. 33 In fact, pseudoscientific ideas arguing Newton's and Einstein's seminal works or those denying climate change that are hardly testable have already found their niche in substandard electronic journals with soft or nonexistent peer review. 34

CITATIONS AND SOCIAL MEDIA ATTENTION

The available preliminary evidence points to the attractiveness of hypothesis articles for readers, particularly those from research-intensive countries who actively download related documents. 35 However, citations of such articles are disproportionately low. Only a small proportion of top-downloaded hypotheses (13%) in the highly prestigious Med Hypotheses receive on average 5 citations per article within a two-year window. 36

With the exception of a few historic papers, the vast majority of hypotheses attract relatively small number of citations in a long term. 36 Plausible explanations are that these articles often contain a single or only a few citable points and that suggested research studies to test hypotheses are rarely conducted and reported, limiting chances of citing and crediting authors of genuine research ideas.

A snapshot analysis of citation activity of hypothesis articles may reveal interest of the global scientific community towards their implications across various disciplines and countries. As a prime example, Strachan's hygiene hypothesis, published in 1989, 10 is still attracting numerous citations on Scopus, the largest bibliographic database. As of August 28, 2019, the number of the linked citations in the database is 3,201. Of the citing articles, 160 are cited at least 160 times ( h -index of this research topic = 160). The first three citations are recorded in 1992 and followed by a rapid annual increase in citation activity and a peak of 212 in 2015 ( Fig. 1 ). The top 5 sources of the citations are Clin Exp Allergy (n = 136), J Allergy Clin Immunol (n = 119), Allergy (n = 81), Pediatr Allergy Immunol (n = 69), and PLOS One (n = 44). The top 5 citing authors are leading experts in pediatrics and allergology Erika von Mutius (Munich, Germany, number of publications with the index citation = 30), Erika Isolauri (Turku, Finland, n = 27), Patrick G Holt (Subiaco, Australia, n = 25), David P. Strachan (London, UK, n = 23), and Bengt Björksten (Stockholm, Sweden, n = 22). The U.S. is the leading country in terms of citation activity with 809 related documents, followed by the UK (n = 494), Germany (n = 314), Australia (n = 211), and the Netherlands (n = 177). The largest proportion of citing documents are articles (n = 1,726, 54%), followed by reviews (n = 950, 29.7%), and book chapters (n = 213, 6.7%). The main subject areas of the citing items are medicine (n = 2,581, 51.7%), immunology and microbiology (n = 1,179, 23.6%), and biochemistry, genetics and molecular biology (n = 415, 8.3%).

An external file that holds a picture, illustration, etc.
Object name is jkms-34-e300-g001.jpg

Interestingly, a recent analysis of 111 publications related to Strachan's hygiene hypothesis, stating that the lack of exposure to infections in early life increases the risk of rhinitis, revealed a selection bias of 5,551 citations on Web of Science. 37 The articles supportive of the hypothesis were cited more than nonsupportive ones (odds ratio adjusted for study design, 2.2; 95% confidence interval, 1.6–3.1). A similar conclusion pointing to a citation bias distorting bibliometrics of hypotheses was reached by an earlier analysis of a citation network linked to the idea that β-amyloid, which is involved in the pathogenesis of Alzheimer disease, is produced by skeletal muscle of patients with inclusion body myositis. 38 The results of both studies are in line with the notion that ‘positive’ citations are more frequent in the field of biomedicine than ‘negative’ ones, and that citations to articles with proven hypotheses are too common. 39

Social media channels are playing an increasingly active role in the generation and evaluation of scientific hypotheses. In fact, publicly discussing research questions on platforms of news outlets, such as Reddit, may shape hypotheses on health-related issues of global importance, such as obesity. 40 Analyzing Twitter comments, researchers may reveal both potentially valuable ideas and unfounded claims that surround groundbreaking research ideas. 41 Social media activities, however, are unevenly distributed across different research topics, journals and countries, and these are not always objective professional reflections of the breakthroughs in science. 2 , 42

Scientific hypotheses are essential for progress in science and advances in healthcare. Innovative ideas should be based on a critical overview of related scientific facts and evidence-based data, often overlooked by others. To generate realistic hypothetical theories, the authors should comprehensively analyze the literature and suggest relevant and ethically sound design for future studies. They should also consider their hypotheses in the context of research and publication ethics norms acceptable for their target journals. The journal editors aiming to diversify their portfolio by maintaining and introducing hypotheses section are in a position to upgrade guidelines for related articles by pointing to general and specific analyses of the subject, preferred study designs to test hypotheses, and ethical implications. The latter is closely related to specifics of hypotheses. For example, editorial recommendations to outline benefits and risks of a new laboratory test or therapy may result in a more balanced article and minimize associated risks afterwards.

Not all scientific hypotheses have immediate positive effects. Some, if not most, are never tested in properly designed research studies and never cited in credible and indexed publication outlets. Hypotheses in specialized scientific fields, particularly those hardly understandable for nonexperts, lose their attractiveness for increasingly interdisciplinary audience. The authors' honest analysis of the benefits and limitations of their hypotheses and concerted efforts of all stakeholders in science communication to initiate public discussion on widely visible platforms and social media may reveal rational points and caveats of the new ideas.

Disclosure: The authors have no potential conflicts of interest to disclose.

Author Contributions:

  • Conceptualization: Gasparyan AY, Yessirkepov M, Kitas GD.
  • Methodology: Gasparyan AY, Mukanova U, Ayvazyan L.
  • Writing - original draft: Gasparyan AY, Ayvazyan L, Yessirkepov M.
  • Writing - review & editing: Gasparyan AY, Yessirkepov M, Mukanova U, Kitas GD.

SEP home page

  • Table of Contents
  • Random Entry
  • Chronological
  • Editorial Information
  • About the SEP
  • Editorial Board
  • How to Cite the SEP
  • Special Characters
  • Advanced Tools
  • Support the SEP
  • PDFs for SEP Friends
  • Make a Donation
  • SEPIA for Libraries
  • Entry Contents

Bibliography

Academic tools.

  • Friends PDF Preview
  • Author and Citation Info
  • Back to Top

The Structure of Scientific Theories

Scientific inquiry has led to immense explanatory and technological successes, partly as a result of the pervasiveness of scientific theories. Relativity theory, evolutionary theory, and plate tectonics were, and continue to be, wildly successful families of theories within physics, biology, and geology. Other powerful theory clusters inhabit comparatively recent disciplines such as cognitive science, climate science, molecular biology, microeconomics, and Geographic Information Science (GIS). Effective scientific theories magnify understanding, help supply legitimate explanations, and assist in formulating predictions. Moving from their knowledge-producing representational functions to their interventional roles (Hacking 1983), theories are integral to building technologies used within consumer, industrial, and scientific milieus.

This entry explores the structure of scientific theories from the perspective of the Syntactic, Semantic, and Pragmatic Views. Each of these answers questions such as the following in unique ways. What is the best characterization of the composition and function of scientific theory? How is theory linked with world? Which philosophical tools can and should be employed in describing and reconstructing scientific theory? Is an understanding of practice and application necessary for a comprehension of the core structure of a scientific theory? Finally, and most generally, how are these three views ultimately related?

1.1 Syntactic, Semantic, and Pragmatic Views: The Basics

1.2 two examples: newtonian mechanics and population genetics, 2.1 theory structure per the syntactic view, 2.2 a running example: newtonian mechanics, 2.3 interpreting theory structure per the syntactic view, 2.4 taking stock: syntactic view, 3.1 theory structure per the semantic view, 3.2 a running example: newtonian mechanics, 3.3 interpreting theory structure per the semantic view, 3.4 taking stock: semantic view, 4.1 theory structure per the pragmatic view, 4.2 a running example: newtonian mechanics, 4.3 interpreting theory structure per the pragmatic view, 4.4 taking stock: pragmatic view, 5. population genetics, 6. conclusion, other internet resources, related entries, 1. introduction.

In philosophy, three families of perspectives on scientific theory are operative: the Syntactic View , the Semantic View , and the Pragmatic View. Savage distills these philosophical perspectives thus:

The syntactic view that a theory is an axiomatized collection of sentences has been challenged by the semantic view that a theory is a collection of nonlinguistic models, and both are challenged by the view that a theory is an amorphous entity consisting perhaps of sentences and models, but just as importantly of exemplars, problems, standards, skills, practices and tendencies. (Savage 1990, vii–viii)

Mormann (2007) characterizes the Syntactic and Semantic Views in similar terms, and is among the first to use the term “Pragmatic View” to capture the third view (137). The three views are baptized via a trichotomy from linguistics deriving from the work of Charles Morris, following Charles S. Peirce. In a classic exposition, the logical positivist Carnap writes:

If in an investigation explicit reference is made to the speaker, or, to put it in more general terms, to the user of a language, then we assign it to the field of pragmatics . (Whether in this case reference to designata is made or not makes no difference for this classification.) If we abstract from the user of the language and analyze only the expressions and their designata, we are in the field of semantics . And if, finally, we abstract from the designata also and analyze only the relations between the expressions, we are in (logical) syntax . The whole science of language, consisting of the three parts mentioned, is called semiotic . (1942, 9; see also Carnap 1939, 3–5, 16)

To summarize, syntax concerns grammar and abstract structures; semantics investigates meaning and representation; and pragmatics explores use. Importantly, while no view is oblivious to the syntax, semantics, or pragmatics of theory, the baptism of each is a product of how one of the three aspects of language is perceived to be dominant: theory as syntactic logical reconstruction (Syntactic View); theory as semantically meaningful mathematical modeling (Semantic View); or theory structure as complex and as closely tied to theory pragmatics, i.e., function and context (Pragmatic View). Each of these philosophical perspectives on scientific theory will be reviewed in this entry. Their relations will be briefly considered in the Conclusion.

It will be helpful to pare each perspective down to its essence. Each endorses a substantive thesis about the structure of scientific theories.

For the Syntactic View, the structure of a scientific theory is its reconstruction in terms of sentences cast in a metamathematical language. Metamathematics is the axiomatic machinery for building clear foundations of mathematics, and includes predicate logic, set theory, and model theory (e.g., Zach 2009; Hacking 2014). A central question of the Syntactic View is: in which logical language should we recast scientific theory?

Some defenders of the Semantic View keep important aspects of this reconstructive agenda, moving the metamathematical apparatus from predicate logic to set theory. Other advocates of the Semantic View insist that the structure of scientific theory is solely mathematical. They argue that we should remain at the mathematical level, rather than move up (or down) a level, into foundations of mathematics. A central question for the Semantic View is: which mathematical models are actually used in science?

Finally, for the Pragmatic View, scientific theory is internally and externally complex. Mathematical components, while often present, are neither necessary nor sufficient for characterizing the core structure of scientific theories. Theory also consists of a rich variety of nonformal components (e.g., analogies and natural kinds). Thus, the Pragmatic View argues, a proper analysis of the grammar (syntax) and meaning (semantics) of theory must pay heed to scientific theory complexity, as well as to the multifarious assumptions, purposes, values, and practices informing theory. A central question the Pragmatic View poses is: which theory components and which modes of theorizing are present in scientific theories found across a variety of disciplines?

In adopting a descriptive perspective on the structure of scientific theories, each view also deploys, at least implicitly, a prescriptive characterization of our central topic. In other words, postulating that scientific theory is \(X\) (e.g., \(X\) = a set-theoretic structure, as per Suppes 1960, 1962, 1967, 1968, 2002) also implies that what is not \(X\) (or could not be recast as \(X\)) is not (or could not possibly be) a scientific theory, and would not help us in providing scientific understanding, explanation, prediction, and intervention. For the Syntactic View, what is not (or cannot be) reconstructed axiomatically is not theoretical, while for the Semantic View, what is not (or cannot be) modeled mathematically is not theoretical. In contrast, in part due to its pluralism about what a scientific theory actually (and possibly) is, and because it interprets theory structure as distributed in practices, the Pragmatic View resists the definitional and normative terms set by the other two views. As a result, the Pragmatic View ultimately reforms the very concepts of “theory” and “theory structure.”

This encyclopedia entry will be organized as follows. After presenting this piece’s two sustained examples, immediately below, the three views are reviewed in as many substantive sections. Each section starts with a brief overview before characterizing that perspective’s account of theory structure. Newtonian mechanics is used as a running example within each section. The interpretation of theory structure—viz., how theory “hooks up” with phenomena, experiment, and the world—is also reviewed in each section. In the final section of this entry, we turn to population genetics and an analysis of the Hardy-Weinberg Principle (HWP) to compare and contrast each view. The Conclusion suggests, and remains non-committal about, three kinds of relations among the views: identity , combat , and complementarity . Theory is not a single, static entity that we are seeing from three different perspectives, as we might represent the Earth using three distinct mathematical map projections. Rather, theory itself changes as a consequence of perspective adopted.

Two examples will be used to illustrate differences between the three views: Newtonian mechanics and population genetics. While relativity theory is the preferred theory of the Syntactic View, Newtonian mechanics is more straightforward. Somewhat permissively construed, the theory of Newtonian mechanics employs the basic conceptual machinery of inertial reference frames, centers of mass, Newton’s laws of motion, etc., to describe the dynamics and kinematics of, among other phenomena, point masses acting vis-à-vis gravitational forces (e.g. the solar system) or with respect to forces involved in collisions (e.g., pool balls on a pool table; a closed container filled with gas). Newtonian mechanics is explored in each section.

Population genetics investigates the genetic composition of populations of natural and domesticated species, including the dynamics and causes of changes in gene frequencies in such populations (for overviews, see Lloyd 1994 [1988]; Gould 2002; Pigliucci and Müller 2010; Okasha 2012). Population genetics emerged as a discipline with the early 20 th century work of R.A. Fisher, Sewall Wright, and J.B.S. Haldane, who synthesized Darwinian evolutionary theory and Mendelian genetics. One important part of population genetic theory is the Hardy-Weinberg Principle. HWP is a null model mathematically stating that gene frequencies remain unchanged across generations when there is no selection, migration, random genetic drift, or other evolutionary forces acting in a given population. HWP peppers early chapters of many introductory textbooks (e.g., Crow and Kimura 1970; Hartl and Clark 1989; Bergstrom and Dugatkin 2012). We return to HWP in Section 5 and here merely state questions each view might ask about population genetics.

The Syntactic View focuses on questions regarding the highest axiomatic level of population genetics (e.g., Williams 1970, 1973; Van Valen 1976; Lewis 1980; Tuomi 1981, 1992). Examples of such queries are:

  • What would be the most convenient metamathematical axiomatization of evolutionary processes (e.g., natural selection, drift, migration, speciation, competition)? In which formal language(s) would and could such axiomatizations be articulated (e.g., first-order predicate logic, set theory, and category theory)?
  • Which single grammars could contain a variety of deep evolutionary principles and concepts, such as HWP, “heritability,” and “competitive exclusion”?
  • Which formal and methodological tools would permit a smooth flow from the metamathematical axiomatization to the mathematical theory of population genetics?

Investigations of the axiomatized rational reconstruction of theory shed light on the power and promises, and weaknesses and incompleteness, of the highest-level theoretical edifice of population genetics.

Secondly, the Semantic View primarily examines questions regarding the mathematical structure of population genetics (Lewontin 1974, Beatty 1981; López Beltrán 1987; Thompson 1989, 2007; Lloyd 1994 [1988]). Very generally, this exploration involves the following questions:

  • What is the form and content of the directly presented class of mathematical models of evolutionary theory (e.g., HWP)? How could and should we organize the cluster of mathematical models (sensu Levins 1966) of population genetics?
  • Which additional models (e.g., diagrammatic, narrative, scale) might be used to enrich our understanding of evolutionary theory?
  • What are the relations among theoretical mathematical models, data models, and experimental models? How does theory explain and shape data? How do the data constrain and confirm theory?

The main subject of investigation is mathematical structure, rather than metamathematics or even alternative model types or modeling methods.

Finally, the Pragmatic View asks about the internal complexity of population genetic theory, as well as about the development and context of population genetics. In so doing, it inquires into how purposes and values have influenced the theoretical structure of evolutionary theory, selecting and shaping current population genetics from a wide variety of possible alternative theoretical articulations. The following questions about the structure of population genetic theory might be here addressed:

  • What role did R.A. Fisher’s interest in animal husbandry, and his tenure at Rothamsted Experimental Station, play in shaping his influential methodologies of Analysis of Variance (ANOVA) and experimental design involving randomization, blocking, and factorial designs?
  • How did the development of computers and computational practices, statistical techniques, and the molecularization of genetics, shape theory and theorizing in population genetics, especially from the 1980s to today?
  • How might normative context surrounding the concept of “race” impact the way concepts such as “heritability” and “lineage,” or principles such as HWP, are deployed in population genetics?

As when studying an organism, the structure of theory cannot be understood independently of its history and function.

2. The Syntactic View

According to the Syntactic View, which emerged mainly out of work of the Vienna Circle and Logical Empiricism (see Coffa 1991; Friedman 1999; Creath 2014; Uebel 2014), philosophy most generally practiced is, and should be, the study of the logic of natural science, or Wissenschaftslogik (Carnap 1937, 1966; Hempel 1966). Robust and clear logical languages allow us to axiomatically reconstruct theories, which—by the Syntacticists’ definition—are sets of sentences in a given logical domain language (e.g., Campbell 1920, 122; Hempel 1958, 46; cf. Carnap 1967 [1928], §156, “Theses about the Constructional System”). Domain languages include “the language of physics, the language of anthropology” (Carnap 1966, 58).

This view has been variously baptized as the Received View (Putnam 1962; Hempel 1970), the Syntactic Approach (van Fraassen 1970, 1989), the Syntactic View (Wessels 1976), the Standard Conception (Hempel 1970), the Orthodox View (Feigl 1970), the Statement View (Moulines 1976, 2002; Stegmüller 1976), the Axiomatic Approach (van Fraassen 1989), and the Once Received View (Craver 2002). For historical reasons, and because of the linguistic trichotomy discussed above, the “Syntactic View” shall be the name of choice in this entry.

Some conceptual taxonomy is required in order to understand the logical framework of the structure of scientific theories for the Syntactic View. We shall distinguish terms , sentences , and languages (see Table 1).

2.1.1 Terms

Building upwards from the bottom, let us start with the three kinds of terms or vocabularies contained in a scientific language: theoretical, logical, and observational. Examples of theoretical terms are “molecule,” “atom,” “proton,” and “protein,” and perhaps even macro-level objects and properties such as “proletariat” and “aggregate demand.” Theoretical terms or concepts can be classificatory (e.g., “cat” or “proton”), comparative (e.g., “warmer”), or quantitative (e.g., “temperature”) (Hempel 1952; Carnap 1966, Chapter 5). Moreover, theoretical terms are “theoretical constructs” introduced “jointly” as a “theoretical system” (Hempel 1952, 32). Logical terms include quantifiers (e.g., \(\forall, \exists\)) and connectives (e.g., \(\wedge, \rightarrow\)). Predicates such as “hard,” “blue,” and “hot,” and relations such as “to the left of” and “smoother than,” are observational terms.

2.1.2 Sentences

Terms can be strung together into three kinds of sentences: theoretical, correspondence, and observational. \(T_S\) is the set of theoretical sentences that are the axioms, theorems, and laws of the theory. Theoretical sentences include the laws of Newtonian mechanics and of the Kinetic Theory of Gases, all suitably axiomatized (e.g., Carnap 1966; Hempel 1966). Primitive theoretical sentences (e.g., axioms) can be distinguished from derivative theoretical sentences (e.g., theorems; see Reichenbach 1969 [1924]; Hempel 1958; Feigl 1970). \(C_S\) is the set of correspondence sentences tying theoretical sentences to observable phenomena or “to a ‘piece of reality’” (Reichenbach 1969 [1924], 8; cf. Einstein 1934, 1936 [1936], 351). To simplify, they provide the theoretical syntax with an interpretation and an application, i.e., a semantics. Suitably axiomatized version of the following sentences provide semantics to Boyle’s law, \(PV = nRT\): “\(V\) in Boyle’s law is equivalent to the measurable volume \(xyz\) of a physical container such as a glass cube that is \(x\), \(y\), and \(z\) centimeters in length, width, and height, and in which the gas measured is contained” and “\(T\) in Boyle’s law is equivalent to the temperature indicated on a reliable thermometer or other relevant measuring device properly calibrated, attached to the physical system, and read.” Carnap (1987 [1932], 466) presents two examples of observational sentences, \(O_S\): “Here (in a laboratory on the surface of the earth) is a pendulum of such and such a kind,” and “the length of the pendulum is 245.3 cm.” Importantly, theoretical sentences can only contain theoretical and logical terms; correspondence sentences involve all three kinds of terms; and observational sentences comprise only logical and observational terms.

2.1.3 Languages

The total domain language of science consists of two languages: the theoretical language, \(L_T\), and the observational language, \(L_O\) (e.g., Hempel 1966, Chapter 6; Carnap 1966, Chapter 23; the index entry for “Language,” of Feigl, Scriven, and Maxwell 1958, 548 has three subheadings: “observation,” “theoretical,” and “ordinary”). The theoretical language includes theoretical vocabulary, while the observational language involves observational terms. Both languages contain logical terms. Finally, the theoretical language includes, and is constrained by, the logical calculus, Calc , of the axiomatic system adopted (e.g., Hempel 1958, 46; Suppe 1977, 50-53). This calculus specifies sentence grammaticality as well as appropriate deductive and non-ampliative inference rules (e.g., modus ponens) pertinent to, especially, theoretical sentences. Calc can itself be written in theoretical sentences.

2.1.4 Theory Structure, in General

Table 1 summarizes the Syntactic View’s account of theory structure:

The salient divide is between theory and observation. Building on Table 1, there are three different levels of scientific knowledge, according to the Syntactic View:

\(\{T_S\} =\) The uninterpreted syntactic system of the scientific theory. \(\{T_S, C_S\} =\) The scientific theory structure of a particular domain (e.g., physics, anthropology). \(\{T_S,C_S,O_S\} =\) All of the science of a particular domain.

Scientific theory is thus taken to be a syntactically formulated set of theoretical sentences (axioms, theorems, and laws) together with their interpretation via correspondence sentences. As we have seen, theoretical sentences and correspondence sentences are cleanly distinct, even if both are included in the structure of a scientific theory.

Open questions remain. Is the observation language a sub-language of the theoretical language, or are they both parts of a fuller language including all the vocabulary? Can the theoretical vocabulary or language be eliminated in favor of a purely observational vocabulary or language? Are there other ways of carving up kinds of languages? First, a “dialectical opposition” between “logic and experience,” “form and content,” “constitutive principles and empirical laws,” and “‘from above’… [and] ‘from below’” pervades the work of the syntacticists (Friedman 1999, 34, 63). Whether syntacticists believe that a synthesis or unification of this general opposition between the theoretical (i.e., logic, form) and the observational (i.e., experience, content) is desirable remains a topic of ongoing discussion. Regarding the second question, Hempel 1958 deflates what he calls “the theoretician’s dilemma”—i.e., the putative reduction without remainder of theoretical concepts and sentences to observational concepts and sentences. Finally, other language divisions are possible, as Carnap 1937 argues (see Friedman 1999, Chapter 7). Returning to the main thread of this section, the distinction toolkit of theoretical and observational terms, sentences, and languages (Table 1) permit the syntacticists to render theoretical structure sharply, thereby aiming at the reconstructive “logic of science” ( Wissenschafstlogik ) that they so desire.

Reichenbach 1969 [1924] stands as a canonical attempt by a central developer of the Syntactic View of axiomatizing a physical theory, viz., relativity theory (cf. Friedman 1983, 1999; see also Reichenbach 1965 [1920]). For the purposes of this encyclopedia entry, it is preferable to turn to another syntactic axiomatization effort. In axiomatizing Newtonian mechanics, the mid-20 th century mathematical logician Hans Hermes spent significant energy defining the concept of mass (Hermes 1938, 1959; Jammer 1961). More precisely, he defines the theoretical concept of “mass ratio” of two particles colliding inelastically in an inertial reference frame \(S\). Here is his full definition of mass ratio (1959, 287):

One paraphrase of this definition is, “‘the mass of \(x\) is α times that of \(x_0\)’ is equivalent to ‘there exists a system \(S\), an instant \(t\), momentary mass points \(y\) and \(y_0\), and initial velocities \(v\) and \(v_0\), such that \(y\) and \(y_0\) are genidentical, respectively, with \(x\) and \(x_0\); the joined mass points move with a velocity of 0 with respect to frame \(S\) immediately upon colliding at time \(t\); and \(y\) and \(y_0\) have determinate velocities \(v\) and \(v_0\) before the collision in the ratio α, which could also be 1 if \(x\) and \(x_0\) are themselves genidentical.’” Hermes employs the notion of “genidentical” to describe the relation between two temporal sections of a given particle’s world line (Jammer 1961, 113). Set aside the worry that two distinct particles cannot be genidentical per Hermes’ definition, though they can have identical properties. In short, this definition is syntactically complete and is written in first-order predicate logic, as are the other axioms and definitions in Hermes (1938, 1959). Correspondence rules connecting a postulated mass \(x\) with an actual mass were not articulated by Hermes.

The link between theory structure and the world, under the Syntactic View, is contained in the theory itself: \(C_S\), the set of correspondence rules. The term “correspondence rules” (Margenau 1950; Nagel 1961, 97–105; Carnap 1966, Chapter 24) has a variety of near-synonyms:

  • Dictionary (Campbell 1920)
  • Operational rules (Bridgman 1927)
  • Coordinative definitions (Reichenbach 1969 [1924], 1938)
  • Reduction sentences (Carnap 1936/1937; Hempel 1952)
  • Correspondence postulates (Carnap 1963)
  • Bridge principles (Hempel 1966; Kitcher 1984)
  • Reduction functions (Schaffner 1969, 1976)
  • Bridge laws (Sarkar 1998)

Important differences among these terms cannot be mapped out here. However, in order to better understand correspondence rules, two of their functions will be considered: (i) theory interpretation (Carnap, Hempel) and (ii) theory reduction (Nagel, Schaffner). The dominant perspective on correspondence rules is that they interpret theoretical terms. Unlike “mathematical theories,” the axiomatic system of physics “cannot have… a splendid isolation from the world” (Carnap 1966, 237). Instead, scientific theories require observational interpretation through correspondence rules. Even so, surplus meaning always remains in the theoretical structure (Hempel 1958, 87; Carnap 1966). Second, correspondence rules are seen as necessary for inter-theoretic reduction (van Riel and Van Gulick 2014). For instance, they connect observation terms such as “temperature” in phenomenological thermodynamics (the reduced theory) to theoretical concepts such as “mean kinetic energy” in statistical mechanics (the reducing theory). Correspondence rules unleash the reducing theory’s epistemic power. Notably, Nagel (1961, Chapter 11; 1979) and Schaffner (1969, 1976, 1993) allow for multiple kinds of correspondence rules, between terms of either vocabulary, in the reducing and the reduced theory (cf. Callender 1999; Winther 2009; Dizadji-Bahmani, Frigg, and Hartmann 2010). Correspondence rules are a core part of the structure of scientific theories and serve as glue between theory and observation.

Finally, while they are not part of the theory structure, and although we saw some examples above, observation sentences are worth briefly reviewing. Correspondence rules attach to the content of observational sentences. Observational sentences were analyzed as (i) protocol sentences or Protokollsätze (e.g., Schlick 1934; Carnap 1987 [1932], 1937, cf. 1963; Neurath 1983 [1932]), and as (ii) experimental laws (e.g., Campbell 1920; Nagel 1961; Carnap 1966; cf. Duhem 1954 [1906]). Although constrained by Calc , the grammar of these sentences is determined primarily by the order of nature, as it were. In general, syntacticists do not consider methods of data acquisition, experiment, and measurement to be philosophically interesting. In contrast, the confirmation relation between (collected) data and theory, especially as developed in inductive logic (e.g., Reichenbach 1938, 1978; Carnap 1962 [1950], 1952), as well as questions about the conventionality, grammaticality, foundationalism, atomism, and content of sense-data and synthetic statements, are considered philosophically important (e.g., Carnap 1987 [1932], 1937, 1966; Neurath 1983 [1932]; Reichenbach 1951; Schlick 1925 [1918], 1934; for contemporary commentary, see, e.g., Creath 1987, 2014; Rutte 1991; Friedman 1999).

To summarize, the Syntactic View holds that there are three kinds of terms or vocabularies: logical, theoretical, and observational; three kinds of sentences: \(T_S\), \(C_S\), and \(O_S\); and two languages: \(L_T\) and \(L_O\). Moreover, the structure of scientific theories could be analyzed using the logical tools of metamathematics. The goal is to reconstruct the logic of science, viz. to articulate an axiomatic system.

Interestingly, this perspective has able and active defenders today, who discuss constitutive and axiomatized principles of the historical “relativized a priori” (Friedman 2001, cf. 2013), argue that “the semantic view, if plausible, is syntactic” (Halvorson 2013), and explore “logicism” for, and in, the philosophy of science (Demopulous 2003, 2013; van Benthem 2012). Furthermore, for purposes of the syntactic reconstruction of scientific theories, some continue espousing—or perhaps plea for the resurrection of—predicate logic (e.g., Lutz 2012, 2014), while other contemporary syntacticists (e.g., Halvorson 2012, 2013, 2019) endorse more recently developed metamathematical and mathematical equipment, such as category theory, which “turns out to be a kind of universal mathematical language like set theory” (Awodey 2006, 2; see Eilenberg and MacLane 1945). Importantly, Halvorson (2019) urges that interlocutors adopt “structured” rather than “flat” views of theories. For the case of the syntactic view this would mean that rather than accept the usual formulation that a theory is a set of sentences, “… [we] might say that a theory consists of both sentences and inferential relations between those sentences” (Halvorson 2019, 277–8). Classical syntacticists such as Rudolf Carnap (Friedman 1999, 2011; Carus 2007; Blatti and Lapointe 2016; Koellner ms. in Other Internet Resources) and Joseph Henry Woodger (Nicholson and Gawne 2014) have recently received increasing attention.

3. The Semantic View

An overarching theme of the Semantic View is that analyzing theory structure requires employing mathematical tools rather than predicate logic. After all, defining scientific concepts within a specific formal language makes any axiomatizing effort dependent on the choice, nature, and idiosyncrasies of that narrowly-defined language. For instance, Suppes understands first-order predicate logic, with its “linguistic” rather than “set-theoretical” entities, as “utterly impractical” for the formalization of “theories with more complicated structures like probability theory” (Suppes 1957, 232, 248–9; cf. Suppes 2002). Van Fraassen, another influential defender of the Semantic View, believes that the logical apparatus of the Syntactic View “had moved us mille milles de toute habitation scientifique , isolated in our own abstract dreams” (van Fraassen 1989, 225). Indeed, what would the appropriate logical language for specific mathematical structures be, especially when such structures could be reconstructed in a variety of formal languages? Why should we imprison mathematics and mathematical scientific theory in syntactically defined language(s) when we could, instead, directly investigate the mathematical objects, relations, and functions of scientific theory?

Consistent with the combat strategy (discussed in the Conclusion), here is a list of grievances against the Syntactic View discussed at length in the work of some semanticists.

  • First-Order Predicate Logic Objection . Theoretical structure is intrinsically and invariably tied to the specific choice of a language, \(L_T\), expressed in first-order predicate logic. This places heavy explanatory and representational responsibility on relatively inflexible and limited languages.
  • Theory Individuation Objection . Since theories are individuated by their linguistic formulations, every change in high-level syntactic formulations will bring forth a distinct theory. This produces a reductio: if \(T_1 = p \rightarrow q\) and \(T_2 = \neg p \vee q\) then \(T_1\) and \(T_2\), though logically equivalent, have different syntactic formulations and would be distinct theories.
  • Theoretical/Observational Languages Objection . Drawing the theoretical/observational distinction in terms of language is inappropriate, as observability pertains to entities rather than to concepts.
  • Unintended Models Objection . There is no clear way of distinguishing between intended and unintended models for syntactically characterized theories (e.g., the Löwenheim-Skolem theorem, Bays 2014).
  • Confused Correspondence Rules Objection . Correspondence rules are a confused medley of direct meaning relationships between terms and world, means of inter-theoretic reduction, causal relationship claims, and manners of theoretical concept testing.
  • Trivially True yet Non-Useful Objection . Presenting scientific theory in a limited axiomatic system, while clearly syntactically correct, is neither useful nor honest, since scientific theories are mathematical structures.
  • Practice and History Ignored Objection . Syntactic approaches do not pay sufficient attention to the actual practice and history of scientific theorizing and experimenting.

What, then, does the Semantic View propose to put in the Syntactic View’s place?

Even a minimal description of the Semantic View must acknowledge two distinct strategies of characterizing and comprehending theory structure: the state-space and the set-/model-theoretic approaches.

3.1.1 The State-Space Approach

The state-space approach emphasizes the mathematical models of actual science, and draws a clear line between mathematics and metamathematics. The structure of a scientific theory is identified with the “class,” “family” or “cluster” of mathematical models constituting it, rather than with any metamathematical axioms “yoked to a particular syntax” (van Fraassen 1989, 366). Under this analysis, “the correct tool for philosophy of science is mathematics, not metamathematics”—this is Suppes’ slogan, per van Fraassen (1989, 221; 1980, 65). In particular, a state space or phase space is an \(N\)-dimensional space, where each of the relevant variables of a theory correspond to a single dimension and each point in that space represents a possible state of a real system. An actual, real system can take on, and change, states according to different kinds of laws, viz., laws of succession determining possible trajectories through that space (e.g., Newtonian kinematic laws); laws of co-existence specifying the permitted regions of the total space (e.g., Boyle’s law); and laws of interaction combining multiple laws of succession or co-existence, or both (e.g., population genetic models combining laws of succession for selection and genetic drift, Wright 1969; Lloyd 1994 [1988]; Rice 2004; Clatterbuck, Sober, and Lewontin 2013). Different models of a given theory will share some dimensions of their state space while differing in others. Such models will also partially overlap in laws (for further discussion of state spaces, laws, and models pertinent to the Semantic View, see Suppe 1977, 224–8; Lloyd 1994, Chapter 2; Nolte 2010; Weisberg 2013, 26–9).

Historically, the state-space approach emerged from work by Evert Beth, John von Neumann, and Hermann Weyl, and has important parallels with Przełęcki (1969) and Dalla Chiara Scabia and Toraldo di Francia (1973) (on the history of the approach see: Suppe 1977; van Fraassen 1980, 65–67; Lorenzano 2013; advocates of the approach include: Beatty 1981; Giere 1988, 2004; Giere, Bickle, and Mauldin 2006; Lloyd 1983, 1994 [1988], 2013 In Press; Suppe 1977, 1989; Thompson, 1989, 2007; van Fraassen 1980, 1989, 2008; for alternative early analyses of models see, e.g., Braithwaite 1962; Hesse 1966, 1967). Interestingly, van Fraassen (1967, 1970) provides a potential reconstruction of state spaces via an analysis of “semi-interpreted languages.” Weisberg (2013), building on many insights from Giere’s work, presents a broad view of modeling that includes mathematical structures that are “trajectories in state spaces” (29), but also permits concrete objects and computational structures such as algorithms to be deemed models. Lorenzano (2013) calls Giere’s (and, by extension, Weisberg’s and even Godfrey-Smith’s 2006) approach “model-based,” separating it out from the state-space approach. A more fine-grained classification of the state-space approach is desirable, particularly if we wish to understand important lessons stemming from the Pragmatic View of Theories, as we shall see below.

As an example of a state-space analysis of modeling, consider a capsule traveling in outer space. An empirically and dynamically adequate mathematical model of the capsule’s behavior would capture the position of the capsule (i.e., three dimensions of the formal state space), as well as the velocity and acceleration vectors for each of the three standard spatial dimensions (i.e., six more dimensions in the formal state space). If the mass were unknown or permitted to vary, we would have to add one more dimension. Possible and actual trajectories of our capsule, with known mass, within this abstract 9-dimensional state space could be inferred via Newtonian dynamical laws of motion (example in Lewontin 1974, 6–8; consult Suppe 1989, 4). Importantly, under the state-space approach, the interesting philosophical work of characterizing theory structure (e.g., as classes of models), theory meaning (e.g., data models mapped to theoretical models), and theory function (e.g., explaining and predicting) happens at the level of mathematical models.

3.1.2 The Set-/Model-Theoretic Approach

Lurking in the background of the state-space conception is the fact that mathematics actually includes set theory and model theory—i.e., mathematical logic. Indeed, according to some interlocutors, “metamathematics is part of mathematics” (Halvorson 2012, 204). Historically, a set-/model-theoretic approach emerged from Tarski’s work and was extensively articulated by Suppes and his associates (van Fraassen 1980, 67). Set theory is a general language for formalizing mathematical structures as collections—i.e., sets—of abstract objects (which can themselves be relations or functions; see Krivine 2013 [1971]). Model theory investigates the relations between, on the one hand, the formal axioms, theorems, and laws of a particular theory and, on the other hand, the mathematical structures—the models—that provide an interpretation of that theory, or put differently, that make the theory’s axioms, theorems, and laws true (Hodges 1997, Chapter 2; Jones 2005). Interestingly, model theory often uses set theory (e.g., Marker 2002); set theory can, in turn, be extended to link axiomatic theories and semantic models via “set-theoretical predicates” (e.g., Suppes 1957, 2002). Finally, there are certain hybrids of these two branches of mathematical logic, including “partial structures” (e.g., da Costa and French 1990, 2003; Bueno 1997; French 2017; French and Ladyman 1999, 2003; Vickers 2009; Bueno, French, and Ladyman 2012). Lorenzano (2013) provides a more complex taxonomy of the intellectual landscape of the Semantic View, including a discussion of Structuralism, a kind of set-/model-theoretic perspective. Structuralism involves theses about “theory-nets,” theory-relative theoretical vs. non-theoretical terms, a diversity of intra- and inter-theoretic laws with different degrees of generality, a typology of inter-theoretic relations, and a rich account of correspondence rules in scientific practice (see Moulines 2002; Pereda 2013; Schmidt 2014; Ladyman 2014). On the whole, the set-/model-theoretic approach of the Semantic View insists on the inseparability of metamathematics and mathematics. In preferring to characterize a theory axiomatically in terms of its intension rather than its extension, it shares the Syntactic View’s aims of reconstructive axiomatization (e.g., Sneed 1979; Stegmüller 1979; Frigg and Votsis 2011; Halvorson 2013, 2019; Lutz 2012, 2014, 2017).

An example will help motivate the relation between theory and model. Two qualifications are required: (i) we return to a more standard set-/model-theoretic illustration below, viz., McKinsey, Sugar, and Suppes’ (1953) axiomatization of particle mechanics, and (ii) this motivational example is not from the heartland of model theory (see Hodges 2013). Following van Fraassen’s intuitive case of “seven-point geometry” (1980, 41–44; 1989, 218–220), also known as “the Fano plane” we see how a particular geometric figure, the model , interprets and makes true a set of axioms and theorems, the theory . In topology and geometry there is rich background theory regarding how to close Euclidean planes and spaces to make finite geometries by, for instance, eliminating parallel lines. Consider the axioms of a projective plane:

  • For any two points, exactly one line lies on both.
  • For any two lines, exactly one point lies on both.
  • There exists a set of four points such that no line has more than two of them.

A figure of a geometric model that makes this theory true is:

Geometric figure including triangle ACE with interior circle BDF and center point G. Point B is on line segment AC, D is on CE, and F is on AE. G is the center of the circle. Point G is on line segments AD, BE, and CF.

This is the smallest geometrical model satisfying the three axioms of the projective plane theory. Indeed, this example fits van Fraassen’s succinct characterization of the theory-model relation:

A model is called a model of a theory exactly if the theory is entirely true if considered with respect to this model alone. (Figuratively: the theory would be true if this model was the whole world.) (1989, 218)

That is, if the entire universe consisted solely of these seven points and seven lines, the projective plane theory would be true. Of course, our universe is bigger. Because Euclidean geometry includes parallel lines, the Fano plane is not a model of Euclidean geometry. Even so, by drawing the plane, we have shown it to be isomorphic to parts of the Euclidean plane. In other words, the Fano plane has been embedded in a Euclidean plane. Below we return to the concepts of embedding and isomorphism, but this example shall suffice for now to indicate how a geometric model can provide a semantics for the axioms of a theory.

In short, for the Semantic View the structure of a scientific theory is its class of mathematical models. According to some advocates of this view, the family of models can itself be axiomatized, with those very models (or other models) serving as axiom truth-makers.

Returning to our running example, consider Suppes’ 1957 model-theoretic articulation of particle mechanics, which builds on his 1953 article with J.C.C. McKinsey and A.C. Sugar. Under this analysis, there is a domain of set-theoretic objects of the form \(\{ P, T, s, m, f, g \}\), where \(P\) and \(T\) are themselves sets, \(s\) and \(g\) are binary functions, \(m\) is a unary and \(f\) a ternary function. \(P\) is the set of particles; \(T\) is a set of real numbers measuring elapsed times; \(s(p, t)\) is the position of particle \(p\) at time \(t\); \(m(p)\) is the mass of particle \(p\); \(f(p, q, t)\) is the force particle \(q\) exerts on \(p\) at time \(t\); and \(g(p, t)\) is the total resultant force (by all other particles) on \(p\) at time \(t\). Suppes and his collaborators defined seven axioms—three kinematical and four dynamical—characterizing Newtonian particle mechanics (see also Simon 1954, 1970). Such axioms include Newton’s third law reconstructed in set-theoretic formulation thus (Suppes 1957, 294):

Importantly, the set-theoretic objects are found in more than one of the axioms of the theory, and Newton’s calculus is reconstructed in a novel, set-theoretic form. Set-theoretic predicates such as “is a binary relation” and “is a function” are also involved in axiomatizing particle mechanics (Suppes 1957, 249). Once these axioms are made explicit, their models can be specified and these can, in turn, be applied to actual systems, thereby providing a semantics for the axioms (e.g., as described in Section 3.3.1 below). A particular system satisfying these seven axioms is a particle mechanics system. (For an example of Newtonian mechanics from the state-space approach, recall the space capsule of Section 3.1.1.)

How is the theory structure, described in Section 3.1, applied to empirical phenomena? How do we connect theory and data via observation and experimental and measuring techniques? The Semantic View distinguishes theory individuation from both theory-phenomena and theory-world relations. Three types of analysis of theory interpretation are worth investigating: (i) a hierarchy of models (e.g., Suppes; Suppe), (ii) similarity (e.g., Giere; Weisberg), and (iii) isomorphism (e.g., van Fraassen; French and Ladyman).

3.3.1 A Hierarchy of Models

One way of analyzing theory structure interpretation is through a series of models falling under the highest-level axiomatizations. This series has been called “a hierarchy of models,” though it need not be considered a nested hierarchy. These models include models of theory, models of experiment, and models of data (Suppes 1962, 2002). Here is a summary of important parts of the hierarchy (Suppes 1962, Table 1, 259; cf. Giere 2010, Figure 1, 270):

  • Axioms of Theory . Axioms define set-theoretic predicates, and constitute the core structure of scientific theories, as reviewed in Section 3.1.2.
  • Models of Theory. “Representation Theorems,” permit us “to discover if an interesting subset of models for the theory may be found such that any model for the theory is isomorphic to some member of this subset” (Suppes 1957, 263). Representation theorem methodology can be extended (i) down the hierarchy, both to models of experiment and models of data, and (ii) from isomorphism to homomorphism (Suppes 2002, p. 57 ff.; Suppe 2000; Cartwright 2008).
  • Models of Experiment . Criteria of experimental design motivate choices for how to set up and analyze experiments. There are complex mappings between models of experiment thus specified, and (i) models of theory, (ii) theories of measurement, and (iii) models of data.
  • Models of Data . In building models of data, phenomena are organized with respect to statistical goodness-of-fit tests and parameter estimation, in the context of models of theory. Choices about which parameters to represent must be made.

The temptation to place phenomena at the bottom of the hierarchy must be resisted because phenomena permeate all levels. Indeed, the “class of phenomena” pertinent to a scientific theory is its “intended scope” (Suppe 1977, 223; Weisberg 2013, 40). Furthermore, this temptation raises fundamental questions about scientific representation: “there is the more profound issue of the relationship between the lower most representation in the hierarchy—the data model perhaps—and reality itself, but of course this is hardly something that the semantic approach alone can be expected to address” (French and Ladyman 1999, 113; cf. van Fraassen 2008, 257–258, “The ‘link’ to reality”). Borrowing from David Chalmers, the “hard problem” of philosophy of science remains connecting abstract structures to concrete phenomena, data, and world.

3.3.2 Similarity

The similarity analysis of theory interpretation combines semantic and pragmatic dimensions (Giere 1988, 2004, 2010; Giere, Bickle, and Mauldin 2006; Weisberg 2013). According to Giere, interpretation is mediated by theoretical hypotheses positing representational relations between a model and relevant parts of the world. Such relations may be stated as follows:

Here \(S\) is a scientist, research group or community, \(W\) is a part of the world, and \(X\) is, broadly speaking, any one of a variety of models (Giere 2004, 743, 747, 2010). Model-world similarity judgments are conventional and intentional:

Note that I am not saying that the model itself represents an aspect of the world because it is similar to that aspect. …Anything is similar to anything else in countless respects, but not anything represents anything else. It is not the model that is doing the representing; it is the scientist using the model who is doing the representing. (2004, 747)

Relatedly, Weisberg (2013) draws upon Tversky (1977) to develop a similarity metric for model interpretation (equation 8.10, 148). This metric combines (i) model-target semantics (90–97), and (ii) the pragmatics of “context, conceptualization of the target, and the theoretical goals of the scientist” (149). Giere and Weisberg thus endorse an abundance of adequate mapping relations between a given model and the world. From this diversity, scientists and scientific communities must select particularly useful similarity relationships for contextual modeling purposes. Because of semantic pluralism and irreducible intentionality, this similarity analysis of theory interpretation cannot be accommodated within a hierarchy of models approach, interpreted as a neat model nesting based on pre-given semantic relations among models at different levels.

3.3.3 Isomorphism

The term “isomorphism” is a composite of the Greek words for “equal” and “shape” or “form.” Indeed, in mathematics, isomorphism is a perfect one-to-one, bijective mapping between two structures or sets. Figure (2) literally and figuratively captures the term:

Script writing of isomorphism with mirror image underneath

Especially in set theory, category theory, algebra, and topology, there are various kinds of “-morphisms,” viz., of mapping relations between two structures or models. Figure (3) indicates five different kinds of homomorphism, arranged in a Venn diagram.

Venn diagram with outer circle Hom and 3 intersecting interior circles: Mon, Epi, and End. The intersection of all 3 is Aut and the intersection of Mon and Epi is Iso.

Although philosophers have focused on isomorphism, other morphisms such as monomorphism (i.e., an injective homomorphism where some elements in the co-domain remain unmapped from the domain) might also be interesting to investigate, especially for embedding data (i.e., the domain) into rich theoretical structures (i.e., the co-domain). To complete the visualization above, an epimorphism is a surjective homomorphism, and an endomorphism is a mapping from a structure to itself, although it need not be a symmetrical—i.e., invertible—mapping, which would be an automorph.

Perhaps the most avid supporter of isomorphism and embedding as the way to understand theory interpretation is van Fraassen. In a nutshell, if we distinguish (i) theoretical models, (ii) “empirical substructures” (van Fraassen 1980, 64, 1989, 227; alternatively: “surface models” 2008, 168), and (iii) “observable phenomena” (1989, 227, 2008, 168), then, van Fraassen argues, theory interpretation is a relation of isomorphism between observable phenomena and empirical substructures, which are themselves isomorphic with one or more theoretical models. Moreover, if a relation of isomorphism holds between \(X\) and a richer \(Y\), we say that we have embedded \(X\) in \(Y\). For instance, with respect to the seven-point geometry above (Figure 1), van Fraassen contends that isomorphism gives embeddability, and that the relation of isomorphism “is important because it is also the exact relation a phenomenon bears to some model or theory, if that theory is empirically adequate” (1989, 219–20; this kind of statement seems to be simultaneously descriptive and prescriptive about scientific representation, see Section 1.1 above). In The Scientific Image he is even clearer about fleshing out the empirical adequacy of a theory (with its theoretical models) in terms of isomorphism between “appearances” (i.e., “the structures which can be described in experimental and measurement reports,” 1980, 64, italics removed) and empirical substructures. Speaking metaphorically,

the phenomena are, from a theoretical point of view, small, arbitrary, and chaotic—even nasty, brutish, and short…—but can be understood as embeddable in beautifully simple but much larger mathematical models. (2008, 247; see also van Fraassen 1981, 666 and 1989, 230)

Interestingly, and as a defender of an identity strategy (see Conclusion), Friedman also appeals to embedding and subsumption relations between theory and phenomena in his analyses of theory interpretation (Friedman 1981, 1983). Bueno, da Costa, French, and Ladyman also employ embedding and (partial) isomorphism in the empirical interpretation of partial structures (Bueno 1997; Bueno, French, and Ladyman 2012; da Costa and French 1990, 2003; French 2017; French and Ladyman 1997, 1999, 2003; Ladyman 2004). Suárez discusses complexities in van Fraassen’s analyses of scientific representation and theory interpretation (Suárez 1999, 2011). On the one hand, representation is structural identity between the theoretical and the empirical. On the other hand, “There is no representation except in the sense that some things are used, made, or taken, to represent some things as thus or so” (van Fraassen 2008, 23, italics removed). The reader interested in learning how van Fraassen simultaneously endorses acontextually structural and contextually pragmatic aspects of representation and interpretation should refer to van Fraassen’s (2008) investigations of maps and “the essential indexical.” [To complement the structure vs. function distinction, see van Fraassen 2008, 309–311 for a structure (“structural relations”) vs. history (“the intellectual processes that lead to those models”) distinction; cf. Ladyman et al. 2011] In all of this, embedding via isomorphism is a clear contender for theory interpretation under the Semantic View.

In short, committing to either a state-space or a set-/model-theoretic view on theory structure does not imply any particular perspective on theory interpretation (e.g., hierarchy of models, similarity, embedding). Instead, commitments to the former are logically and actually separable from positions on the latter (e.g., Suppes and Suppe endorse different accounts of theory structure, but share an understanding of theory interpretation in terms of a hierarchy of models). The Semantic View is alive and well as a family of analyses of theory structure, and continues to be developed in interesting ways both in its state-space and set-/model-theoretic approaches.

4. The Pragmatic View

The Pragmatic View recognizes that a number of assumptions about scientific theory seem to be shared by the Syntactic and Semantic Views. Both perspectives agree, very roughly, that theory is (1) explicit, (2) mathematical, (3) abstract, (4) systematic, (5) readily individualizable, (6) distinct from data and experiment, and (7) highly explanatory and predictive (see Flyvbjerg 2001, 38–39; cf. Dreyfus 1986). The Pragmatic View imagines the structure of scientific theories rather differently, arguing for a variety of theses:

  • Limitations . Idealized theory structure might be too weak to ground the predictive and explanatory work syntacticists and semanticists expect of it (e.g., Cartwright 1983, 1999a, b, 2019; Morgan and Morrison 1999; Suárez and Cartwright 2008).
  • Pluralism . Theory structure is plural and complex both in the sense of internal variegation and of existing in many types. In other words, there is an internal pluralism of theory (and model) components (e.g., mathematical concepts, metaphors, analogies, ontological assumptions, values, natural kinds and classifications, distinctions, and policy views, e.g., Kuhn 1970; Boumans 1999), as well as a broad external pluralism of different types of theory (and models) operative in science (e.g., mechanistic, historical, and mathematical models, e.g., Hacking 2009, Longino 2013). Indeed, it may be better to speak of the structures of scientific theories, in the double-plural.
  • Nonformal aspects. The internal pluralism of theory structure (thesis #2) includes many nonformal aspects deserving attention. That is, many components of theory structure, such as metaphors, analogies, values, and policy views have a non-mathematical and “informal” nature, and they lie implicit or hidden (e.g., Bailer-Jones 2002; Craver 2002; Contessa 2006; Morgan 2012). Interestingly, the common understanding of “formal,” which identifies formalization with mathematization, may itself be a conceptual straightjacket; the term could be broadened to include “diagram abstraction” and “principle extraction” (e.g., Griesemer 2013, who explicitly endorses what he also calls a “Pragmatic View of Theories”).
  • Function. Characterizations of the nature and dynamics of theory structure should pay attention to the user as well as to purposes and values (e.g., Apostel 1960; Minsky 1965; Morrison 2007; Winther 2012a).
  • Practice . Theory structure is continuous with practice and “the experimental life,” making it difficult to neatly dichotomize theory and practice (e.g., Hacking 1983, 2009; Shapin and Schaffer 1985; Galison 1987, 1988, 1997; Suárez and Cartwright 2008, Cartwright 2019).

These are core commitments of the Pragmatic View.

It is important to note at the outset that the Pragmatic View takes its name from the linguistic trichotomy discussed above, in the Introduction. This perspective need not imply commitment to, or association with, American Pragmatism (e.g. the work of Charles S. Peirce, William James, or John Dewey; cf. Hookway 2013; Richardson 2002). For instance, Hacking (2007a) distinguishes his pragmatic attitudes from the school of Pragmatism. He maps out alternative historical routes of influence, in general and on him, vis-à-vis fallibilism (via Imre Lakatos, Karl Popper; Hacking 2007a, §1), historically conditioned truthfulness (via Bernard Williams; Hacking 2007a, §3), and realism as intervening (via Francis Everitt, Melissa Franklin; Hacking 2007a, §4). To borrow a term from phylogenetics, the Pragmatic View is “polyphyletic.” The components of its analytical framework have multiple, independent origins, some of which circumnavigate American Pragmatism.

With this qualification and the five theses above in mind, let us now turn to the Pragmatic View’s analysis of theory structure and theory interpretation.

We should distinguish two strands of the Pragmatic View: the Pragmatic View of Models and a proper Pragmatic View of Theories .

4.1.1 The Pragmatic View of Models

Nancy Cartwright’s How the Laws of Physics Lie crystallized the Pragmatic View of Models. Under Cartwright’s analysis, models are the appropriate level of investigation for philosophers trying to understand science. She argues for significant limitations of theory (thesis #1), claiming that laws of nature are rarely true, and are epistemically weak. Theory as a collection of laws cannot, therefore, support the many kinds of inferences and explanations that we have come to expect it to license. Cartwright urges us to turn to models and modeling, which are central to scientific practice. Moreover, models “lie”—figuratively and literally—between theory and the world (cf. Derman 2011). That is, “to explain a phenomenon is to find a model that fits it into the basic framework of the theory and that thus allows us to derive analogues for the messy and complicated phenomenological laws which are true of it.” A plurality of models exist, and models “serve a variety of purposes” (Cartwright 1983, 152; cf. Suppes 1978). Cartwright is interested in the practices and purposes of scientific models, and asks us to focus on models rather than theories.

Cartwright’s insights into model pluralism and model practices stand as a significant contribution of “The Stanford School” (cf. Cat 2014), and were further developed by the “models as mediators” group, with participants at LSE, University of Amsterdam, and University of Toronto (Morgan and Morrison 1999; Chang 2011; cf. Martínez 2003). This group insisted on the internal pluralism of model components (thesis #2). According to Morgan and Morrison, building a model involves “fitting together… bits which come from disparate sources,” including “stories” (Morgan and Morrison 1999, 15). Boumans (1999) writes:

model building is like baking a cake without a recipe. The ingredients are theoretical ideas, policy views, mathematisations of the cycle, metaphors and empirical facts. (67) Mathematical moulding is shaping the ingredients in such a mathematical form that integration is possible… (90)

In an instructive diagram, Boumans suggests that a variety of factors besides theory and data feed into a model: metaphors, analogies, policy views, stylised facts, mathematical techniques, and mathematical concepts (93). The full range of components involved in a model will likely vary according to discipline, and with respect to explanations and interventions sought (e.g., analogies but not policy views will be important in theoretical physics). In short, model building involves a complex variety of internal nonformal aspects, some of which are implicit (theses #2 and #3).

As one example of a nonformal component of model construction and model structure, consider metaphors and analogies (e.g., Bailer-Jones 2002). Geary (2011) states the “simplest equation” of metaphor thus: “\(X = Y\)” (8, following Aristotle: “Metaphor consists in giving the thing a name that belongs to something else… ,” Poetics , 1457b). The line between metaphor and analogy in science is blurry. Some interlocutors synonymize them (e.g., Hoffman 1980; Brown 2003), others reduce one to the other (analogy is a form of metaphor, Geary 2011; metaphor is a kind of analogy, Gentner 1982, 2003), and yet others bracket one to focus on the other (e.g., Oppenheimer 1956 sets aside metaphor). One way to distinguish them is to reserve “analogy” for concrete comparisons, with clearly identifiable and demarcated source and target domains, and with specific histories, and use “metaphor” for much broader and indeterminate comparisons, with diffuse trajectories across discourses. Analogies include the “lines of force” of electricity and magnetism (Maxwell and Faraday), the atom as a planetary system (Rutherford and Bohr), the benzene ring as a snake biting its own tail (Kekulé), Darwin’s “natural selection” and “entangled bank,” and behavioral “drives” (Tinbergen) (e.g., Hesse 1966, 1967; Bartha 2010). Examples of metaphor are genetic information, superorganism, and networks (e.g., Keller 1995). More could be said about other informal model components, but this discussion of metaphors and analogies shall suffice to hint at how models do not merely lie between theory and world. Models express a rich internal pluralism (see also de Chadarevian and Hopwood 2004; Morgan 2012).

Model complexity can also be seen in the external plurality of models (thesis #2). Not all models are mathematical, or even ideally recast as mathematical. Non-formalized (i.e., non–state-space, non-set-/model-theoretic) models such as physical, diagrammatic, material, historical, “remnant,” and fictional models are ubiquitous across the sciences (e.g., Frigg and Hartmann 2012; for the biological sciences, see Hull 1975; Beatty 1980; Griesemer 1990, 1991 a, b, 2013; Downes 1992; Richards 1992; Winther 2006a; Leonelli 2008; Weisberg 2013). Moreover, computer simulations differ in important respects from more standard analytical mathematical models (e.g., Smith 1996; Winsberg 2010; Weisberg 2013). According to some (e.g., Griesemer 2013; Downes 1992; Godfrey-Smith 2006; Thomson-Jones 2012), this diversity belies claims by semanticists that models can always be cast “into set theoretic terms” (Lloyd 2013 In Press), are “always a mathematical structure” (van Fraassen 1970, 327), or that “formalisation of a theory is an abstract representation of the theory expressed in a formal deductive framework… in first-order predicate logic with identity, in set theory, in matrix algebra and indeed, any branch of mathematics...” (Thompson 2007, 485–6). Even so, internal pluralism has been interpreted as supporting a “deflationary semantic view,” which is minimally committed to the perspective that “model construction is an important part of scientific theorizing” (Downes 1992, 151). Given the formal and mathematical framework of the Semantic View (see above), however, the broad plurality of kinds of models seems to properly belong under a Pragmatic View of Models.

4.1.2 The Pragmatic View of Theories

Interestingly, while critiquing the Syntactic and Semantic Views on most matters, the Pragmatic View of Models construed theory, the process of theorizing, and the structure of scientific theories, according to terms set by the two earlier views. For instance, Cartwright tends to conceive of theory as explicit, mathematical, abstract, and so forth (see the first paragraph of Section 4). She always resisted “the traditional syntactic/semantic view of theory” for its “vending machine” view, in which a theory is a deductive and automated machine that upon receiving empirical input “gurgitates” and then “drops out the sought-for representation” (1999a, 184–5). Rather than reform Syntactic and Semantic accounts of theory and theory structure, however, she invites us, as we just saw, to think of science as modeling, “with theory as one small component” (Cartwright, Shomar, and Suárez 1995, 138; Suárez and Cartwright 2008). Many have followed her. Kitcher’s predilection is also to accept the terms of the Syntactic and Semantic Views. For instance, he defines theories as “axiomatic deductive systems” (1993, 93). In a strategy complementary to Cartwright’s modeling turn, Kitcher encourages us to focus on practice, including practices of modeling and even practices of theorizing. In The Advancement of Science , practice is analyzed as a 7-tuple, with the following highly abbreviated components: (i) a language; (ii) questions; (iii) statements (pictures, diagrams); (iv) explanatory patterns; (v) standard examples; (vi) paradigms of experimentation and observation, plus instruments and tools; and (vii) methodology (Kitcher 1993, 74). Scientific practice is also center stage for those singing the praises of “the experimental life” (e.g., Hacking 1983; Shapin and Schaffer 1985; Galison 1987), and those highlighting the cognitive grounds of science (e.g., Giere 1988; Martínez 2014) and science’s social and normative context (e.g., Kitcher 1993, 2001; Longino 1995, 2002; Ziman 2000; cf. Simon 1957). Indeed, the modeling and practice turns in the philosophy of science were reasonable reactions to the power of axiomatic reconstructive and mathematical modeling analyses of the structure of scientific theories.

Yet, a Pragmatic View of Theories is also afoot, one resisting orthodox characterizations of theory often embraced, at least early on, by Pragmatic View philosophers such as Cartwright, Hacking, Kitcher, and Longino. For instance, Craver (2002) accepts both the Syntactic and Semantic Views, which he humorously and not inaccurately calls “the Once Received View” and the “Model Model View.” But he also observes:

While these analyses have advanced our understanding of some formal aspects of theories and their uses, they have neglected or obscured those aspects dependent upon nonformal patterns in theories. Progress can be made in understanding scientific theories by attending to their diverse nonformal patterns and by identifying the axes along which such patterns might differ from one another. (55)

Craver then turns to mechanistic theory as a third theory type (and a third philosophical analysis of theory structure) that highlights nonformal patterns:

Different types of mechanisms can be distinguished on the basis of recurrent patterns in their organization. Mechanisms may be organized in series, in parallel, or in cycles. They may contain branches and joins, and they often include feedback and feedforward subcomponents. (71)

Consistent with theses #2 and #3 of the Pragmatic View, we must recognize the internal pluralism of theories as including nonformal components. Some of these are used to represent organizational and compositional relations of complex systems (Craver 2007; Wimsatt 2007; Winther 2011; Walsh 2015). While mechanistic analyses such as Craver’s may not wish to follow every aspect of the Pragmatic View of Theories, there are important and deep resonances between the two.

In a review of da Costa and French (2003), Contessa (2006) writes:

Philosophers of science are increasingly realizing that the differences between the syntactic and the semantic view are less significant than semanticists would have it and that, ultimately, neither is a suitable framework within which to think about scientific theories and models. The crucial divide in philosophy of science, I think, is not the one between advocates of the syntactic view and advocates of the semantic view, but the one between those who think that philosophy of science needs a formal framework or other and those who think otherwise. (376)

Again, we are invited to develop a non-formal framework of science and presumably also of scientific theory. (Halvorson 2012, 203 takes Contessa 2006 to task for advocating “informal philosophy of science.”) Moreover, in asking “what should the content of a given theory be taken to be on a given occasion?”, Vickers (2009) answers:

It seems clear that, in addition to theories being vague objects in the way that ‘heaps’ of sand are, there will be fundamentally different ways to put together theoretical assumptions depending on the particular investigation one is undertaking. For example, sometimes it will be more appropriate to focus on the assumptions which were used by scientists, rather than the ones that were believed to be true. (247, footnote suppressed)

A Pragmatic View of Theories helps make explicit nonformal internal components of theory structure.

Key early defenders of the modeling and practice turns have also recently begun to envision theory in a way distinct from the terms set by the Syntactic and Semantic Views. Suárez and Cartwright (2008) extend and distribute theory by arguing that “What we know ‘theoretically’ is recorded in a vast number of places in a vast number of different ways—not just in words and formulae but in machines, techniques, experiments and applications as well” (79). And while her influence lies primarily in the modeling turn, even in characterizing the “vending machine” view, Cartwright calls for a “reasonable philosophical account of theories” that is “much more textured, and… much more laborious” than that adopted by the Syntactic and Semantic Views (1999a, 185). The theory-data and theory-world axes need to be rethought. In her 2019 book on “artful modeling”, Cartwright emphasizes the importance of know-how and creativity in scientific practice, and “praise[s] engineers and cooks and inventors, as well as experimental physicists like Millikan and Melissa Franklin” (Cartwright 2019, 76). Kitcher wishes to transform talk of theories into discussion of “significance graphs” (2001, 78 ff.). These are network diagrams illustrating which (and how) questions are considered significant in the context of particular scientific communities and norms (cf. Brown 2010). Consistently with a Pragmatic View of Theories, Morrison (2007) reconsiders and reforms canonical conceptualizations of “theory.” Finally, Longino (2013) proposes an archaeology of assumptions behind and under different research programs and theories of human behavior such as neurobiological, molecular behavioral genetic, and social-environmental approaches (e.g., Oyama 2000). For instance, two shared or recurring assumptions across programs and theories are:

(1) that the approach in question has methods of measuring both the behavioral outcome that is the object of investigation and the factors whose association with it are the topic of investigation and (2) that the resulting measurements are exportable beyond the confines of the approach within which they are made. (Longino 2013, 117)

A Pragmatic View of Theories expands the notion of theory to include nonformal aspects, which surely must include elements from Boumans’ list above (e.g., metaphors, analogies, policy views), as well as more standard components such as ontological assumptions (e.g., Kuhn 1970; Levins and Lewontin 1985; Winther 2006b), natural kinds (e.g., Hacking 2007b), and conditions of application or scope (e.g., Longino 2013).

In addition to exploring internal theory diversity and in parallel with plurality of modeling, a Pragmatic View of Theories could also explore pluralism of modes of theorizing, and of philosophically analyzing theoretical structure (thesis #2). Craver (2002) provides a start in this direction in that he accepts three kinds of scientific theory and of philosophical analysis of scientific theory. A more synoptic view of the broader pragmatic context in which theories are embedded can be found in the literature on different “styles” of scientific reasoning and theorizing (e.g., Crombie 1994, 1996; Vicedo 1995; Pickstone 2000; Davidson 2001; Hacking 2002, 2009; Winther 2012b; Elwick 2007; Mancosu 2010). While there is no univocal or dominant classification of styles, two lessons are important. First, a rough consensus exists that theoretical investigations of especially historical, mechanistic, and mathematical structures and relations will involve different styles. Second, each style integrates theoretical products and theorizing processes in unique ways, thus inviting an irreducible pragmatic methodological pluralism in our philosophical analysis of the structure of scientific theories. For instance, the structure of theories of mechanisms in molecular biology or neuroscience involves flow charts, and is distinct from the structure of theories of historical processes and patterns as found in systematics and phylogenetics, which involves phylogenetic trees. As Crombie suggests, we need a “comparative historical anthropology of thinking.” (1996, 71; see Hacking 2009) Mathematical theory hardly remains regnant. It gives way to a pluralism of theory forms and theory processes. Indeed, even mathematical theorizing is a pluralistic motley, as Hacking (2014) argues. Although a “deflationary” Semantic View could account for pluralism of theory forms, the Pragmatic View of Theories, drawing on styles, is required to do justice to the immense variety of theorizing processes, and of philosophical accounts of theory and theory structure.

Finally, outstanding work remains in sorting out the philosophical utility of a variety of proposed units in addition to styles, such as Kuhn’s (1970) paradigms, Lakatos’ (1980) research programmes, Laudan’s (1977) research traditions, and Holton’s (1988) themata. A rational comparative historical anthropology of both theorizing and philosophical analyses of theorizing remains mostly unmapped (cf. Matheson and Dallmann 2014). Such a comparative meta-philosophical analysis should also address Davidson’s (1974) worries about “conceptual schemes” and Popper’s (1996 [1976]) critique of “the myth of the framework” (see Hacking 2002; Godfrey-Smith 2003).

Cartwright has done much to develop a Pragmatic View. Start by considering Newton’s second law:

Here \(F\) is the resultant force on a mass \(m\), and \(a\) is the net acceleration of \(m\); both \(F\) and \(a\) are vectors. This law is considered a “general” (Cartwright 1999a, 187) law expressed with “abstract quantities” (Cartwright 1999b, 249). Newton’s second law can be complemented with other laws, such as (i) Hooke’s law for an ideal spring:

Here \(k\) is the force constant of the spring, and \(x\) the distance along the x-axis from the equilibrium position, and (ii) Coulomb’s law modeling the force between two charged particles:

Here \(K\) is Coulomb’s electrical constant, \(q\) and \(q'\) are the charges of the two objects, and \(r\) the distance between the two objects. The picture Cartwright draws for us is that Newton’s, Hooke’s, and Coulomb’s laws are abstract, leaving out many details. They can be used to derive mathematical models of concrete systems. For instance, by combining (1) and (2), the law of gravitation (a “fundamental” law, Cartwright 1983, 58–59), other source laws, and various simplifying assumptions, we might create a model for the orbit of Mars, treating the Sun and Mars as a 2-body system, ignoring the other planets, asteroids, and Mars’ moons. Indeed, the Solar System is a powerful “nomological machine” (Cartwright 1999a, 50–53), which “is a fixed (enough) arrangement of components, or factors, with stable (enough) capacities that in the right sort of stable (enough) environment will, with repeated operation, give rise to the kind of regular behaviour that we represent in our scientific laws” (Cartwright 1999a, 50). Importantly, most natural systems are complex and irregular, and cannot be neatly characterized as nomological machines. For these cases, abstract laws “run out” (Cartwright 1983) and are rarely smoothly “deidealised” (Suárez 1999). In general, abstract laws predict and explain only within a given domain of application, and only under ideal conditions. More concrete laws or models are not directly deduced from them (e.g., Suárez 1999, Suárez and Cartwright 2008), and they can rarely be combined to form effective “super-laws” (Cartwright 1983, 70–73). In short, the move from (1) and (2) or from (1) and (3) to appropriate phenomenological models, is not fully specified by either abstract law pairing. Indeed, Cartwright developed her notion of “capacities” to discuss how “the principles of physics” “are far better rendered as claims about capacities, capacities that can be assembled and reassembled in different nomological machines, unending in their variety, to give rise to different laws” (1999a, 52). Articulating concrete models requires integrating a mix of mathematical and nonformal components. Laws (1), (2), and (3) remain only one component, among many, of the models useful for, e.g., exploring the behavior of the Solar System, balls on a pool table, or the behavior of charges in electrical fields.

Shifting examples but not philosophical research program, Suárez and Cartwright (2008) explains how analogies such as superconductors as diamagnets (as opposed to ferromagnets) were an integral part of the mathematical model of superconductivity developed by Fritz and Heinz London in the 1930s (63; cf. London and London 1935). Suárez and Cartwright gladly accept that this model “is uncontroversially grounded in classic electromagnetic theory” (64). However, contra Semantic View Structuralists such as Bueno, da Costa, French, and Ladyman, they view nonformal aspects as essential to practices of scientific modeling and theorizing: “The analogy [of diamagnets] helps us to understand how the Londons work with their model… which assumptions they add and which not… a formal reconstruction of the model on its own cannot help us to understand that” (69). In short, the running example of Newtonian mechanics, in conjunction with a glimpse into the use of analogies in mathematical modeling, illustrates the Pragmatic View’s account of theory syntax: theory is constituted by a plurality of formal and informal components.

As we have explored throughout this section, models and theories have informal internal components, and there are distinct modes of modeling and theorizing. Because of the Pragmatic View’s attention to practice, function, and application, distinguishing structure from interpretation is more difficult here than under the Syntactic and Semantic Views. Any synchronic analysis of the structure of models and theories must respect intentional diachronic processes of interpreting and using, as we shall now see.

Regarding the import of function in models and theories (thesis #4), already the Belgian philosopher of science Apostel defined modeling thus: “Let then \(R(S,P,M,T)\) indicate the main variables of the modelling relationship. The subject \(S\) takes, in view of the purpose \(P\), the entity \(M\) as a model for the prototype \(T\)” (1960, 128, see also Apostel 1970). Purposes took center-stage in his article title: “Towards the Formal Study of Models in the Non-Formal Sciences.” MIT Artificial Intelligence trailblazer Minsky also provided a pragmatic analysis:

We use the term “model” in the following sense: To an observer \(B\), an object \(A^*\) is a model of an object \(A\) to the extent that \(B\) can use \(A^*\) to answer questions that interest him about \(A\). The model relation is inherently ternary. Any attempt to suppress the role of the intentions of the investigator \(B\) leads to circular definitions or to ambiguities about “essential features” and the like. (1965, 45)

This account is thoroughly intentionalist and anti-essentialist. That is, mapping relations between model and world are left open and overdetermined. Specifying the relevant relations depends on contextual factors such as questions asked, and the kinds of similarities and isomorphisms deemed to be of interest. The appropriate relations are selected from an infinite (or, at least, near-infinite) variety of possible relations (e.g., Rosenblueth and Wiener 1945; Lowry 1965).

Regarding practice (thesis #5), in addition to ample work on the experimental life mentioned above, consider a small example. A full understanding of the content and structure of the London brothers’ model of superconductivity requires attention to informal aspects such as analogies. Even London and London (1935) state in the summary of their paper that “the current [”in a supraconductor“] is characterized as a kind of diamagnetic volume current” (88). They too saw the diamagnetic analogy as central to their theoretical practices. Criteria and practices of theory confirmation also differ from the ones typical of the Syntactic and Semantic Views. While predictive and explanatory power as well as empirical adequacy remain important, the Pragmatic View also insists on a variety of other justificatory criteria, including pragmatic virtues (sensu Kuhn 1977; Longino 1995) such as fruitfulness and utility. In a nutshell, the Pragmatic View argues that scientific theory structure is deeply shaped and constrained by functions and practices, and that theory can be interpreted and applied validly according to many different criteria.

The analytical framework of the Pragmatic View remains under construction. The emphasis is on internal diversity, and on the external pluralism of models and theories, of modeling and theorizing, and of philosophical analyses of scientific theories. The Pragmatic View acknowledges that scientists use and need different kinds of theories for a variety of purposes. There is no one-size-fits-all structure of scientific theories. Notably, although the Pragmatic View does not necessarily endorse the views of the tradition of American Pragmatism, it has important resonances with the latter school’s emphasis on truth and knowledge as processual, purposive, pluralist, and context-dependent, and on the social and cognitive structure of scientific inquiry.

A further qualification in addition to the one above regarding American Pragmatism is in order. The Pragmatic View has important precursors in the historicist or “world view” perspectives of Feyerabend, Hanson, Kuhn, and Toulmin, which were an influential set of critiques of the Syntactic View utterly distinct from the Semantic View. This philosophical tradition focused on themes such as meaning change and incommensurability of terms across world views (e.g., paradigms), scientific change (e.g., revolutionary: Kuhn 1970; evolutionary: Toulmin 1972), the interweaving of context of discovery and context of justification, and scientific rationality (Preston 2012; Bird 2013; Swoyer 2014). The historicists also opposed the idea that theories can secure meaning and empirical support from a theory-neutral and purely observational source, as the Syntactic View had insisted on with its strong distinction between theoretical and observational vocabularies (cf. Galison 1988). Kuhn’s paradigms or, more precisely, “disciplinary matrices” even had an internal anatomy with four components: (i) laws or symbolic generalizations, (ii) ontological assumptions, (iii) values, and (iv) exemplars (Kuhn 1970, postscript; Godfrey-Smith 2003; Hacking 2012). This work was concerned more with theory change than with theory structure and had fewer conceptual resources from sociology of science and history of science than contemporary Pragmatic View work. Moreover, paradigms never quite caught on the way analyses of models and modeling have. Even so, this work did much to convince later scholars, including many of the Pragmatic View, of certain weaknesses in understanding theories as deductive axiomatic structures.

As a final way to contrast the three views, we return to population genetics and, especially, to the Hardy-Weinberg Principle (HWP). Both Woodger (1937, 1959) and Williams (1970, 1973) provide detailed axiomatizations of certain parts of biology, especially genetics, developmental biology, and phylogenetics. For instance, Woodger (1937) constructs an axiomatic system based on ten logical predicates or relations, including \(\bP\) ( part of ), \(\bT\) ( before in time ), \(\bU\) ( reproduced by cell division or cell fusion ), \(\bm\) ( male gamete ), \(\bff\) ( female gamete ), and \(\bgenet\) ( genetic property ) (cf. Nicholson and Gawne 2014). Woodger (1959) elaborates these logical predicates or relations to produce a careful reconstruction of Mendelian genetics. Here are two axioms in his system (which are rewritten in contemporary notation, since Woodger used Russell and Whitehead’s Principia Mathematica notation):

The first axiom should be read thus: “no gamete is both male and female” (1959, 416). In the second axiom, given that \(DLZxyz\) is a primitive relation defined as “\(x\) is a zygote which develops in the environment \(y\) into the life \(z\)” (1959, 415), the translation is “every life develops in one and only one environment from one and only one zygote” (416). Woodger claims that “the whole of Mendel’s work can be expressed…” via this axiomatic system. Woodger briefly mentions that if one assumes that the entire system or population is random with respect to gamete fusions, “then the Pearson-Hardy law is derivable” (1959, 427). This was a reference to HWP. In her explorations of various axiomatizations of Darwinian lineages and “subclans,” and the process of the “expansion of the fitter,” Williams (1970, 1973) also carefully defines concepts, and axiomatizes basic biological principles of reproduction, natural selection, fitness, and so forth. However, she does not address HWP. Of interest is the lack of axiomatization of HWP or other mathematical principles of population genetics in Woodger’s and Williams’ work. Were such principles considered secondary or uninteresting by Woodger and Williams? Might Woodger’s and Williams’ respective axiomatic systems simply lack the power and conceptual resources to axiomatically reconstruct a mathematical edifice actually cast in terms of probability theory? Finally, other friends of the Syntactic View, such as the early Michael Ruse, do not provide an axiomatization of HWP (Ruse 1975, 241).

Proponents of the Semantic View claim that their perspective on scientific theory accurately portrays the theoretical structure of population genetics. Thompson (2007) provides both set-theoretical and state-space renditions of Mendelian genetics. The first involves defining a set-theoretic predicate for the system, viz., \(\{P, A, f, g\}\), where \(P\) and \(A\) are sets representing, respectively, the total collection of alleles and loci in the population, while \(f\) and \(g\) are functions assigning an allele to a specific location in, respectively, the diploid cells of an individual or the haploid gametic cells. Axioms in this set-theoretic formalization include “The sets \(P\) and \(A\) are finite and non empty” (2007, 498). In contrast, the state-space approach of the Semantic View articulates a phase space with each dimension representing allelic (or genotypic) frequencies (e.g., cover and Chapter 3 of Lloyd 1994 [1988]). As an example, “for population genetic theory, a central law of succession is the Hardy-Weinberg law” (Thompson 2007, 499). Mathematically, the diploid version of HWP is written thus:

Here \(p\) and \(q\) are the frequencies of two distinct alleles at a biallelic locus. The left-hand side represents the allele frequencies in the parental generation and a random mating pattern, while the right-hand side captures genotype frequencies in the offspring generation, as predicted from the parental generation. This is a null theoretical model—actual genotypic and allelic frequencies of the offspring generation often deviate from predicted frequencies (e.g., a lethal homozygote recessive would make the \(q^2_{\text{off}}\) term = 0). Indeed, HWP holds strictly only in abstracted and idealized populations with very specific properties (e.g., infinitely large, individuals reproduce randomly) and only when there are no evolutionary forces operating in the population (e.g., no selection, mutation, migration, or drift) (e.g., Hartl and Clark 1989; Winther et al. 2015). HWP is useful also in the way it interacts with laws of succession for selection, mutation, and so forth (e.g., Okasha 2012). This powerful population genetic principle is central to Semantic View analyses of the mathematical articulation of the theoretical structure of population genetics (see also Lorenzano 2014, Ginnobili 2016).

Recall that the Pragmatic View highlights the internal and external pluralism—as well as the purposiveness—of model and theory structure. Consider recent uses of population genetic theory to specify the kinds and amounts of population structure existing in Homo sapiens . In particular, different measures and mathematical modeling methodologies are employed in investigating human genomic diversity (e.g., Jobling et al. 2004; Barbujani et al. 2013; Kaplan and Winther 2013). It is possible to distinguish at least two different research projects, each of which has a unique pragmatic content (e.g., aims, values, and methods). Diversity partitioning assesses genetic variation within and among pre-determined groups using Analysis of Variance (also crucial to estimating heritability, Downes 2014). Clustering analysis uses Bayesian modeling techniques to simultaneously produce clusters and assign individuals to these “unsupervised” cluster classifications. The robust result of the first modeling project is that (approximately) 85% of all genetic variance is found within human subpopulations (e.g., Han Chinese or Sami), 10% across subpopulations within a continental region, and only 5% is found across continents (i.e., “African,” “Asian,” and “European” – Lewontin 1972, 1974). (Recall also that we are all already identical at, on average, 999 out of 1000 nucleotides.) To calculate diversity partitions at these three nested levels, Lewontin (1972) used a Shannon information-theoretic measure closely related to Sewall Wright’s \(F\)-statistic:

Here \(H_T\) is the total heterozygosity of the population assessed, and \(\bar{H}_S\) is the heterozygosity of each subpopulation (group) of the relevant population, averaged across all the subpopulations. \(F_{ST}\) is bounded by 0 and 1, and is a measure of population structure, with higher \(F_{ST}\) values suggesting more structure, viz., more group differentiation. HWP appears implicitly in both \(H_T\) and \(\bar{H}_S\), which take heterozygosity (\(2pq\)) to be equal to the expected proportion of heterozygotes under HWP rather than the actual frequency of heterozygotes. \(H_T\) is computed by using the grand population average of \(p\) and \(q\), whereas calculating \(\bar{H}_S\) involves averaging across the expected heterozygosities of each subpopulation. If random mating occurs—and thus HWP applies—across the entire population without respecting subpopulation borders, then \(H_T\) and \(\bar{H}_S\) will be equal (i.e., \(p\) of the total population and of each individual subpopulation will be the same; likewise for \(q\)). If, instead, HWP applies only within subpopulations but not across the population as a whole, then \(\bar{H}_S\) will be smaller than \(H_T\), and \(F_{ST}\) will be positive (i.e., there will be “excess homozygosity” across subpopulations, which is known as the “Wahlund Principle” in population genetics). This is one way among many to deploy the population-genetic principle of HWP. Thus, the Lewontin-style diversity partitioning result that only roughly 5% of the total genetic variance is among races is equivalent to saying that \(F_{ST}\) across the big three continental populations in Lewontin’s three-level model is 0.05 (e.g., Barbujani et al. 1997). The basic philosophical tendency is to associate the diversity partitioning research project’s (approximately) 85%-10%-5% result with an anti-realist interpretation of biological race.

In contrast, clustering analysis (e.g., Pritchard et al. 2000; Rosenberg et al. 2002; cf. Edwards 2003) can be readily performed even with the small amount of among-continent genetic variance in Homo sapiens . For instance, when the Bayesian modeling computer program STRUCTURE is asked to produce 5 clusters, continental “races” appear—African, Amerindian, Asian, European, and Pacific Islanders. Interestingly, this modeling technique is also intimately linked to HWP: “Our main modeling assumptions are Hardy-Weinberg equilibrium within populations and complete linkage equilibrium between loci within populations” (Pritchard et al. 2000, 946). That is, for a cluster to eventually be robust in the modeling runs, it should meet HWP expectations. Clustering analysis has sometimes been interpreted as a justification for a realist stance towards biological race (see discussions in Hochman 2013; Winther and Kaplan 2013; Edge and Rosenberg 2015; Spencer 2015).

This example of the mathematical modeling of human genomic diversity teaches that basic and simple formal components can be used in different ways to develop and apply theory, both inside and outside of science. In contrast to the Syntactic and Semantic Views, the Pragmatic View foregrounds tensions vis-à-vis ontological assumptions and political consequences regarding the existence (or not) of biological race between diversity partitioning (Lewontin 1972) and clustering analysis (Pritchard et al. 2000) research packages. These ontological ruptures can be identified despite the fact that both research projects assess population structure by examining departures from HWP (i.e., they measure excess homozygosity), and are completely consistent (e.g., Winther 2014; Ludwig 2015; Edge and Rosenberg 2015).

This exploration of how the three views on the structure of scientific theory address population genetics, and in particular HWP, invites a certain meta-pluralism. That is, the Syntactic View carefully breaks down fundamental concepts and principles in genetics and population genetics, articulating definitions and relations among terms. The Semantic View insightfully decomposes and interweaves the complex mathematical edifice of population genetics. The Pragmatic View sheds light on modeling choices and on distinct interpretations and applications of the same theory or model, both within and without science. The three perspectives are hardly mutually exclusive. (N.B., the two running examples concern theory structure in Newtonian mechanics and population genetics, independently considered. While interesting, debates about “evolutionary forces” are beyond the scope of the current entry; see, e.g., Hitchcock and Velasco 2014.)

The structure of scientific theories is a rich topic. Theorizing and modeling are core activities across the sciences, whether old (e.g., relativity theory, evolutionary theory) or new (e.g., climate modeling, cognitive science, and systems biology). Furthermore, theory remains essential to developing multipurpose tools such as statistical models and procedures (e.g., Bayesian models for data analysis, agent-based models for simulation, network theory for systems analysis). Given the strength and relevance of theory and theorizing to the natural sciences, and even to the social sciences (e.g., microeconomics, physical, if not cultural, anthropology), philosophical attention to the structure of scientific theories could and should increase. This piece has focused on a comparison of three major perspectives: Syntactic View, Semantic View, and Pragmatic View. In order to handle these complex debates effectively, we have sidestepped certain key philosophical questions, including questions about scientific realism; scientific explanation and prediction; theoretical and ontological reductionism; knowledge-production and epistemic inference; the distinction between science and technology; and the relationship between science and society. Each of these topics bears further philosophical investigation in light of the three perspectives here explored.

A table helps summarize general aspects of the three views’ analyses of the structure of scientific theories:

Table 2. General aspects of each view’s analysis of the structure of scientific theories.

The Syntactic, Semantic, and Pragmatic views are often taken to be mutually exclusive and, thus, to be in competition with one another. They indeed make distinct claims about the anatomy of scientific theories. But one can also imagine them to be complementary, focusing on different aspects and questions of the structure of scientific theories and the process of scientific theorizing. For instance, in exploring nonformal and implicit components of theory, the Pragmatic View accepts that scientific theories often include mathematical parts, but tends to be less interested in these components. Moreover, there is overlap in questions—e.g., Syntactic and Semantic Views share an interest in formalizing theory; the Semantic and Pragmatic Views both exhibit concern for scientific practice.

How are these three views ultimately related? A standard philosophical move is to generalize and abstract, understanding a situation from a higher level. One “meta” hypothesis is that a given philosophical analysis of theory structure tends to be associated with a perceived relationship among the three views here discussed. The Syntactic View is inclined to interpret the Semantic View’s formal machinery as continuous with its own generalizing axiomatic strategy, and hence diagnoses many standard Semantic View critiques (Section 3) as missing their mark (the strategy of identity ; e.g., Friedman 1982; Worrall 1984; Halvorson 2012, 2013, 2019; Lutz 2012, 2017; cf. Chakravartty 2001). The Semantic View explicitly contrasts its characterization of theory structure with the “linguistic” or “metamathematical” apparatus of the Syntactic View (the strategy of combat ; e.g., Suppe 1977; van Fraassen 1980, 1989; Lloyd 1994 [1988]). Finally, the Pragmatic View, which did not exist as a perspective until relatively recently, imagines theory as pluralistic and can thus ground a holistic philosophical investigation. It envisions a meta-pluralism in which reconstructive axiomatization and mathematical modeling remain important, though not necessary for all theories. This third view endorses a panoply of theoretical structures and theorizing styles, negotiating continuity both between theorizing and “the experimental life,” and among philosophical analyses of the structure of scientific theories (the strategy of complementarity ; e.g., Hacking 1983, 2009; Galison 1988, 1997; Craver 2002; Suárez and Cartwright 2008; Griesemer 2013). Interestingly, Suárez and Pero (2019) explicitly concur with the Pragmatic View as described in this article, but believe that “the semantic conception in its bare minimal expression” is compatible with, if not sufficient for, capturing “pragmatic elements and themes involved in a more flexible and open-ended approach to scientific theory” (Suárez and Pero 2019, 348). By design, the ecumenical meta-pluralism sanctioned by the Pragmatic View does not completely offset identity and combat strategies. Moreover, only “partial acceptance” of the respective views may ultimately be possible. Even so, the complementarity strategy might be worth developing further. Compared to identity and combat meta-perspectives, it provides broader—or at least different—insights into the structure of scientific theories. More generally, exploring the relations among these views is itself a rich topic for future philosophical work, as is investigating their role in, and interpretation of, active scientific fields ripe for further philosophical analysis such as climate change (e.g., Winsberg 2018), model organisms (e.g., Ankeny and Leonelli 2020), and cartography and GIS (e.g., Winther 2020).

  • Ankeny, R. and S. Leonelli, 2020, Model Organisms , Cambridge: Cambridge University Press.
  • Apostel, L., 1960, “Towards the Formal Study of Models in the Non-Formal Sciences,” Synthese , 12 (23): 125–161.
  • –––, 1970, “The Justification of Formalisation,” Quality and Quantity , 4 (1): 3–38.
  • Awodey, S., 2006, Category Theory , Oxford: Oxford University Press.
  • Bailer-Jones, D.M., 2002, “Models, Metaphors and Analogies,” in Blackwell Guide to the Philosophy of Science , P.K. Machamer and M. Silberstein (eds.), Oxford: Blackwell, pp. 108–127.
  • Barbujani, G., S. Ghirotto, and F. Tassi, 2013, “Nine Things to Remember about Human Genome Diversity,” Tissue Antigens , 82 (3): 155–164.
  • Barbujani, G.A., Magagni, E. Minch, and L.L. Cavalli-Sforza, 1997, “An Apportionment of Human DNA Diversity,” Proceedings of the National Academy of Sciences , 94 (9): 4516–4519.
  • Bartha, P.F.A., 2010, By Parallel Reasoning: The Construction and Evaluation of Analogical Arguments , New York: Oxford University Press
  • Bays, T., 2014, “Skolem’s Paradox”, The Stanford Encyclopedia of Philosophy (Spring 2014 Edition), E. N. Zalta (ed.), URL = < https://plato.stanford.edu/archives/spr2014/entries/paradox-skolem/ >.
  • Beatty, J., 1981, “What’s Wrong with the Received View of Evolutionary Theory?” PSA: Proceedings of the Biennial Meeting of the Philosophy of Science Association 1980 , (2): 397–426.
  • Bergstrom, C. and L. Dugatkin, 2012, Evolution , New York: Norton.
  • Bird, A., 2013, “Thomas Kuhn”, The Stanford Encyclopedia of Philosophy (Fall 2013 Edition), E. N. Zalta (ed.), URL = < https://plato.stanford.edu/archives/fall2013/entries/thomas-kuhn/ >.
  • Blatti, S. and S. Lapointe (eds.), 2016, Ontology After Carnap , Oxford: Oxford University Press.
  • Boumans, M., 1999, “Built-In Justification,” in Models as Mediators: Perspectives on Natural and Social Science , M.S. Morgan and M. Morrison (eds.), Cambridge: Cambridge University Press, pp. 66–96.
  • Braithwaite, R., 1962, “Models in the Empirical Sciences,” in Logic, Methodology and Philosophy of Science: Proceedings of the 1960 International Congress , E. Nagel, P. Suppes, and A. Tarski (eds.), Stanford, CA: Stanford University Press, pp. 224–231.
  • Bridgman, P.W., 1927, The Logic of Modern Physics , New York: Macmillan.
  • Bueno, O., 1997, “Empirical Adequacy: A Partial Structures Approach,” Studies in History and Philosophy of Science (Part A) , 28 (4): 585–610.
  • Bueno, O., S. French, and J. Ladyman, 2012, “Models and Structures: Phenomenological and Partial,” Studies in History and Philosophy of Science (Part B) , 43 (1): 43–46.
  • Brown, T., 2003, Making Truth: Metaphor in Science , Urbana: University of Illinois Press.
  • Brown, M.J., 2010, “Genuine Problems and the Significance of Science,” Contemporary Pragmatism , 7 (2): 131–153.
  • Callender, C., 1999, “Reducing Thermodynamics to Statistical Mechanics: The Case of Entropy,” The Journal of Philosophy , 96 (7): 348–373.
  • Campbell, N.R., 1920, Physics: The Elements , Cambridge: Cambridge University Press.
  • Carnap, R., 1967 [1928], The Logical Structure of the World , translated by R.A. George, Berkeley, CA: University of California Press. Original: Der logische Aufbau der Welt , Leipzig: Felix Meiner.
  • –––, 1932, “Über Protokollsätze”, Erkenntnis , 3: 215–228; transl. by R. Creath and R. Nollan, “On Protocol Sentences,” Noûs , 21 (4) (1987): 457–470.
  • –––, 1936/1937, “Testability and Meaning,” Philosophy of Science , 1936, 3 (4): 419–471; 1937, 4 (1): 1–40.
  • –––, 1937, The Logical Syntax of Language , London: Kegan Paul, Trench, & Trübner.
  • –––, 1939, Foundations of Logic and Mathematics (International Encyclopedia of Unified Science, Volume 1, Number 3), Chicago: University of Chicago Press.
  • –––, 1942, Introduction to Semantics , Cambridge, MA: Harvard University Press.
  • –––, 1952, The Continuum of Inductive Methods , Chicago: University of Chicago Press.
  • –––, 1962 [1950], Logical Foundations of Probability , Chicago: University of Chicago Press, 2 nd edition.
  • –––, 1963, “Philosopher Replies,” in The Philosophy of Rudolf Carnap (Library of Living Philosophers, Volume 11), P. Schilpp (ed.), La Salle: Open Court, pp. 889–999.
  • –––, 1966, Philosophical Foundations of Science , New York: Basic Books; repr. as An Introduction to the Philosophy of Science , 1972; repr. New York: Dover, 1996.
  • Cartwright, N., 1983, How the Laws of Physics Lie , New York: Oxford University Press.
  • –––, 1989, Nature’s Capacities and Their Measurement , New York: Oxford University Press.
  • –––, 1999a, The Dappled World: A Study of the Boundaries of Science , Cambridge: Cambridge University Press.
  • –––, 1999b, “Models and the Limits of Theories: Quantum Hamiltonians and the BCS Model of Superconductivity,” in Models as Mediators: Perspectives on Natural and Social Science , M. Morgan and M. Morrison (eds.), (Perspectives on Natural and Social Sciences), Cambridge: Cambridge University Press, pp. 241–281.
  • –––, 2008, “In Praise of the Representation Theorem,” in Representation, Evidence, and Justification: Themes from Suppes , W.K. Essler and M. Frauchiger (eds.), Ontos Verlag, pp. 83–90.
  • –––, 2019, Nature, the Artful Modeler: Lectures on Laws, Science, How Nature Arranges the World and How We Can Arrange It Better , Chicago, IL: Open Court.
  • Cartwright, N., T. Shomar, and M. Suárez, 1995, “The Tool Box of Science: Tools for the Building of Models with a Superconductivity Example,” in Theories and Models in Scientific Processes (Poznan Studies in the Philosophy of the Sciences and the Humanities, Volume 44), W. Herfel, W. Krajewski, I. Niiniluoto, and R. Wojcicki (eds.), Amsterdam: Rodopi, pp. 137–149.
  • Carus, A.W., 2007, Carnap and Twentieth-Century Thought: Explication as Enlightenment , Cambridge: Cambridge University Press.
  • Cat, J., 2014, “The Unity of Science”, The Stanford Encyclopedia of Philosophy (Winter 2014 Edition), E. N. Zalta (ed.), URL = < https://plato.stanford.edu/archives/win2014/entries/scientific-unity/ >.
  • Chakravartty, A., 2001, “The Semantic or Model-Theoretic View of Theories and Scientific Realism,” Synthese , 127 (3): 325–345.
  • Chang, H., 2011, “The Philosophical Grammar of Scientific Practice” in International Studies in the Philosophy of Science , 25 (3): 205–221.
  • Clatterbuck, H., E. Sober, and R. Lewontin, 2013, “Selection Never Dominates Drift (Nor Vice Versa),” Biology & Philosophy , 28 (4): 577–592.
  • Coffa, A. J., 1991, The Semantic Tradition From Kant to Carnap: To the Vienna Station , Cambridge: Cambridge University Press.
  • Contessa, G., 2006, “Scientific Models, Partial Structures and the New Received View of Theories,” Studies in History and Philosophy of Science (Part A) , 37 (2): 370–377.
  • Craver, C.F., 2002, “Structures of Scientific Theories,” in Blackwell Guide to the Philosophy of Science , P.K. Machamer and M. Silberstein (eds.), Oxford: Blackwell, pp. 55–79.
  • –––, 2007, Explaining the Brain: Mechanisms and the Mosaic Unity of Neuroscience , New York: Oxford University Press.
  • Creath, R., 1987, “The Initial Reception of Carnap’s Doctrine of Analyticity,” Noûs , 21 (4): 477–499.
  • –––, 2014, “Logical Empiricism”, The Stanford Encyclopedia of Philosophy (Spring 2014 Edition), E. N. Zalta (ed.), URL = < https://plato.stanford.edu/archives/spr2014/entries/logical-empiricism/ >.
  • Crombie, A.C., 1994, Styles of Scientific Thinking in the European Tradition (Volumes 1–3), London: Duckworth.
  • –––, 1996, “Commitments and Styles of European Scientific Thinking,” Theoria , 11 (25): 65–76.
  • Crow J. and M. Kimura, 1970, An Introduction to Population Genetics Theory , Edina, MN: Burgess International Group Incorporated.
  • da Costa, N.C.A. and S. French, 1990, “The Model-Theoretic Approach in the Philosophy of Science,” Philosophy of Science , 57 (2): 248–65.
  • –––, 2003. Science and Partial Truth: A Unitary Approach to Models and Scientific Reasoning , Oxford: Oxford University Press.
  • Dalla Chiara Scabia, M.L. and G. Toraldo di Francia, 1973, “A Logical Analysis of Physical Theories,” La Rivista del Nuovo Cimento , 3 (1): 1–20.
  • Davidson, A., 2001, The emergence of sexuality: Historical epistemology and the formation of concepts , Cambridge, MA: Harvard University Press.
  • Davidson, D., 1974, “On the Very Idea of a Conceptual Scheme,” Proceedings and Addresses of the American Philosophical Association , 47: 5–20.
  • de Chadarevian, S. and N. Hopwood, 2004, Models: The Third Dimension of Science , Stanford, CA: Stanford University Press.
  • Demopoulos, W., 2003, “On the Rational Reconstruction of our Theoretical Knowledge,” The British Journal for the Philosophy of Science , 54 (3): 371–403.
  • –––, 2013, Logicism and Its Philosophical Legacy , Cambridge: Cambridge University Press.
  • Derman, E., 2011, Models Behaving Badly: Why Confusing Illusion with Reality Can Lead to Disaster, on Wall Street and in Life , New York: Free Press.
  • Dizadji-Bahmani, F., R. Frigg, and S. Hartmann, 2010, “Who’s Afraid of Nagelian Reduction?,” Erkenntnis , 73 (3): 393–412.
  • Döring, A. and R.G. Winther, forthcoming, “The Human Condition is an Ocean: Philosophy and the Mediterranean Sea,” in Words and Worlds: Use and Abuse of Analogies and Metaphors within Sciences and Humanities , S. Wuppuluri and A.C. Grayling (eds.), Synthese Library Series.
  • Downes, S., 1992, “The Importance of Models in Theorizing: A Deflationary Semantic View,” PSA: Proceedings of the Biennial Meeting of the Philosophy of Science Association 1992 , (1): 142–153.
  • –––, “Heritability,” The Stanford Encyclopedia of Philosophy (Spring 2014 Edition), E. N. Zalta (ed.), URL = < https://plato.stanford.edu/archives/spr2014/entries/heredity/ >.
  • Dreyfus, H., 1986, “Why Studies of Human Capacities Modeled on Ideal Natural Science Can Never Achieve their Goal,” in Rationality, Relativism, and the Human Sciences , J. Margolis, M. Krausz, and R. Burian (eds.), Dordrecht: Martinus Nijhoff, pp. 3–22.
  • Duhem, P., 1906, La théorie physique: Son objet et sa structure , Paris: Chevalier et Rivière; transl. by P.W. Wiener, The Aim and Structure of Physical Theory , Princeton, NJ: Princeton University Press (1954).
  • Edge, M.D. and N. Rosenberg, 2015, “Implications of the Apportionment of Human Genetic Diversity for the Apportionment of Human Phenotypic Diversity,” Studies in History and Philosophy of Biological and Biomedical Sciences , 52: 32–45.
  • Edwards, A.W.F., 2003, “Human Genetic Diversity: Lewontin‘s Fallacy” BioEssays , 25 (8): 798–801.
  • Eilenberg, S. and S. MacLane, 1945, “General Theory of Natural Equivalences,” Transactions of the American Mathematical Society , 58 (2): 231–294.
  • Einstein, A., 1934, “On the Method of Theoretical Physics,” Philosophy of Science , 1 (2): 163–169.
  • –––, 1936, “Physik und Realität,” Journal of The Franklin Institute , 221 (3): 313–347; transl. by J. Piccard, “Physics and Reality,” Journal of the Franklin Institute , 221 (3) (1936): 349–382.
  • Elwick, J., 2007, Styles of Reasoning in British Life Sciences: Shared Assumptions, 1820–1858 , London: Pickering & Chatto.
  • Feigl, H., 1970, “The ‘Orthodox’ View of Theories: Remarks in Defense as Well as Critique,” in Analyses of Theories and Methods of Physics and Psychology (Minnesota Studies in the Philosophy of Science, Volume 4), M. Radner and S. Winokur (eds.), Minneapolis: University of Minnesota Press, pp. 3–16.
  • Feigl, H., M. Scriven, and G. Maxwell (eds.), 1958, Minnesota Studies in the Philosophy of Science (Volume 2), Minneapolis: University of Minnesota Press.
  • Flyvbjerg, B., 2001, Making Social Science Matter: Why Social Inquiry Fails and How it Can Succeed Again , Cambridge: Cambridge University Press.
  • French, S., 2017, “Identity Conditions, Idealisations and Isomorphisms: a Defence of the Semantic Approach,” first online 19 September 2017, Synthese . doi:10.1007/s11229-017-1564-z
  • French, S. and J. Ladyman, 1997, “Superconductivity and Structures: Revisiting the London Account,” Studies in History and Philosophy of Modern Physics , 28 (3): 363–393.
  • –––, 1999, “Reinflating the Semantic Approach,” International Studies in the Philosophy of Science , 13 (2): 103–121.
  • –––, 2003. “Remodelling Structural Realism: Quantum Physics and the Metaphysics of Structure,” Synthese , 136 (1): 31–56.
  • Friedman, M., 1981, “Theoretical Explanation,” in Reduction, Time, and Reality: Studies in the Philosophy of the Natural Sciences , R. Healey (ed.), New York: Cambridge University Press, pp. 1–16.
  • –––, 1982, “ The Scientific Image , by B. van Fraassen,” The Journal of Philosophy , 79 (5): 274–283.
  • –––, 1983, Foundations of Space-Time Theories: Relativistic Physics and Philosophy of Science , Princeton: Princeton University Press.
  • –––, 1999, Reconsidering Logical Positivism , New York: Cambridge University Press.
  • –––, 2001, Dynamics of Reason , Stanford, CA: CSLI Publications.
  • –––, 2011, “Carnap on Theoretical Terms: Structuralism without Metaphysics,” Synthese , 180 (2): 249–263.
  • –––, 2013, Kant’s Construction of Nature: A Reading of the Metaphysical Foundations of Natural Science , Cambridge: Cambridge University Press.
  • Frigg, R. and S. Hartmann, 2012, “Models in Science”, The Stanford Encyclopedia of Philosophy (Fall 2012 Edition), E. N. Zalta (ed.), URL = < https://plato.stanford.edu/archives/fall2012/entries/models-science/ >.
  • Frigg, R. and I. Votsis, 2011, “Everything You Always Wanted to Know about Structural Realism but Were Afraid to Ask,” European Journal for Philosophy of Science , 1 (2): 227–276.
  • Galison, P., 1987, How Experiments End , Chicago: University of Chicago Press.
  • –––, 1988, “History, Philosophy, and the Central Metaphor,” Science in Context , 2 (1): 197–212.
  • –––, 1997, Image and Logic: A Material Culture of Microphysics , Chicago: University of Chicago Press.
  • Geary, J., 2011, I Is an Other: The Secret Life of Metaphor and How It Shapes the Way We See The World , New York: Harper Perennial.
  • Gentner, D., 1982, “Are Scientific Analogies Metaphors?” in Metaphor: Problems and Perspectives , D. Miall (ed.), Brighton: Harvester Press, pp. 106–132.
  • –––, 2003, “Analogical Reasoning, Psychology of,” in Encyclopedia of Cognitive Science , L. Nadel (ed.), London: Nature Publishing Group, pp. 106–112.
  • Giere, R., 1988, Explaining Science: A Cognitive Approach , Chicago: University of Chicago Press.
  • –––, 2004, “How Models Are Used to Represent Reality,” Philosophy of Science , 71 (5): 742–752.
  • –––, 2010, “An Agent-based Conception of Models and Scientific Representation,” Synthese , 172 (2): 269–281.
  • Giere, R., B. Bickle, and R. Mauldin, 2006, Understanding Scientific Reasoning , Belmont, CA: Thomson/Wadsworth, 5 th edition.
  • Ginnobili, S., 2016, “Missing Concepts in Natural Selection Theory Reconstructions,” History and Philosophy of the Life Sciences , 38 (Article 8). doi:10.1007/s40656-016-0109-y
  • Godfrey-Smith, P., 2003, Theory and Reality: An Introduction to the Philosophy of Science , Chicago: University of Chicago Press.
  • –––, 2006, “The Strategy of Model-Based Science,” Biology and Philosophy , 21 (5): 725–740.
  • Gould, S.J., 2002, The Structure of Evolutionary Theory , Cambridge, MA: Harvard University Press.
  • Griesemer, J., 1990, “Modeling in the Museum: On the Role of Remnant Models in the Work of Joseph Grinnell,” Biology and Philosophy , 5 (1): 3–36.
  • –––, 1991a, “Material Models in Biology,” PSA: Proceedings of the Biennial Meeting of the Philosophy of Science Association 1990 , (2): 79–94.
  • –––, 1991b, “Must Scientific Diagrams Be Eliminable? The Case of Path Analysis,” Biology and Philosophy , 6 (2): 155–180.
  • –––, 2013, “Formalization and the Meaning of Theory in the Inexact Biological Sciences,” Biological Theory , 7 (4): 298–310.
  • Hacking, I., 1983, Representing and Intervening: Introductory Topics in the Philosophy of Natural Science , Cambridge: Cambridge University Press.
  • –––, 2002, Historical Ontology , Cambridge, MA: Harvard University Press.
  • –––, 2007a, “On Not Being a Pragmatist: Eight Reasons and a Cause,” in New Pragmatists , C. Misak (ed.), New York: Oxford University Press, pp. 32–49.
  • –––, 2007b, “Natural Kinds: Rosy Dawn, Scholastic Twilight,” Royal Institute of Philosophy Supplements , 61: 203–240.
  • –––, 2009, Scientific Reason , Taipei: National Taiwan University Press.
  • –––, 2012, “Introduction,” in T.S. Kuhn, The Structure of Scientific Revolutions , 50 th Anniversary ed. (4 th ed.), Chicago: University of Chicago Press, pp. vii–xxxvii.
  • –––, 2014, Why Is There Philosophy of Mathematics At All? , Cambridge: Cambridge University Press.
  • Halvorson, H., 2012, “What Scientific Theories Could Not Be,” Philosophy of Science , 79 (2): 183–206.
  • –––, 2013, “The Semantic View, if Plausible, is Syntactic,” Philosophy of Science , 80 (3): 475–478.
  • –––, 2019, The Logic in Philosophy of Science , Cambridge: Cambridge University Press.
  • Hartl, D. and A. Clark, 1989, Principles of Population Genetics , Sunderland, MA: Sinauer Associates.
  • Hempel, C., 1952, Fundamentals of Concept Formation in Empirical Science , Chicago: University of Chicago Press.
  • –––, 1958, “The Theoretician’s Dilemma,” in Minnesota Studies in the Philosophy of Science (Volume 2), H. Feigl, M. Scriven, and G. Maxwell (eds.), Minneapolis: University of Minnesota Press, pp. 37–98.
  • –––, 1966, Philosophy of Natural Science , Englewood Cliffs, N.J.: Prentice-Hall.
  • –––, 1970, “On the ‘Standard Conception’ of Scientific Theories,” in Minnesota Studies in the Philosophy of Science (Volume 4), M. Radner and S. Winokur (eds.), Minneapolis: University of Minnesota Press, pp. 142–163.
  • Hermes, H. 1938, Eine Axiomatisierung der allgemeinen Mechanik (Forschungen zur Logik und zur Grundlegung der exacten Wissenschaften, Heft 3), Leipzig: S. Hirzel.
  • –––, 1959, “Zur Axiomatisierung der Mechanik,” in The Axiomatic Method with Special Reference to Geometry and Physics: Proceedings of an International Symposium Held at the University of California, Berkeley, December 26, 1957–January 4, 1958 , L. Henkin, P. Suppes, and A. Tarski (eds.), Amsterdam: North Holland, pp. 282–290.
  • Hesse, M., 1966, Models and Analogies in Science , Notre Dame: University of Notre Dame Press.
  • –––, 1967, “Models and Analogy in Science,” in The Encyclopedia of Philosophy (Volume 5), P. Edwards (ed.), New York: Macmillan, pp. 354–359.
  • Hitchcock, C. and J.D. Velasco, 2014, “Newtonian and Evolutionary Forces,” Ergo , 1 (2): 39–77.
  • Hochman, A., 2013, “Against the New Racial Naturalism,” The Journal of Philosophy 110 (6): 331–351.
  • Hodges, W., 1997, A Shorter Model Theory , New York: Cambridge University Press.
  • –––, 2013, “Model Theory”, The Stanford Encyclopedia of Philosophy (Fall 2013 Edition), E. N. Zalta (ed.), URL = < https://plato.stanford.edu/archives/fall2013/entries/model-theory/ >.
  • Hoffman, R., 1980, “Metaphor in Science,” in Cognition and Figurative Language , R. Honeck (ed.), Hillsdale: Lawrence Erlbaum Associates, pp. 393–423.
  • Holton, G., 1988, Thematic Origins of Scientific Thought: Kepler to Einstein , Cambridge, MA: Harvard University Press, 2 nd edition.
  • Hookway, C., 2013, “Pragmatism”, The Stanford Encyclopedia of Philosophy (Winter 2013 Edition), E. N. Zalta (ed.), URL = < https://plato.stanford.edu/archives/win2013/entries/pragmatism/ >.
  • Hull, D., 1975, “Central Subjects and Historical Narratives,” History & Theory , 14 (3): 253–274.
  • Jammer, M., 1961, Concepts of Mass in Classical and Modern Physics , Cambridge, MA: Harvard University Press; reprinted unabridged by Dover in 1997.
  • Jobling, M.A., M. Hurles, C. Tyler-Smith, 2004, Human Evolutionary Genetics. Origins, Peoples and Diseases , New York: Garland Science.
  • Jones, M., 2005, “Idealization and Abstraction: A Framework,” in Idealization XII: Correcting the Model – Idealization and Abstraction in the Sciences (Poznan Studies in the Philosophy of the Sciences and the Humanities, Volume 86), M. Jones and N. Cartwright (eds.), Amsterdam: Rodopi, pp. 173–217. (Same individual as Thomson-Jones 2012.)
  • Kaplan, J.M. and R.G. Winther, 2013, “Prisoners of Abstraction? The Theory and Measure of Genetic Variation, and the Very Concept of ‘Race’,” Biological Theory , 7 (4): 401–412.
  • Keller, E.F., 1995, Reconfiguring Life: Metaphors of Twentieth-Century Biology , New York: Columbia University Press.
  • Kitcher P., 1984, “1953 and All That. A Tale of Two Sciences,” Philosophical Review , 93 (3): 335–373.
  • –––, 1993, The Advancement of Science: Science Without Legend, Objectivity Without Illusion , New York: Oxford University Press.
  • –––, 2001, Science, Truth, and Democracy , New York: Oxford University Press.
  • Krivine, J., 2013 [1971], Introduction to Axiomatic Set Theory (Synthese Library, Volume 34), Dordrecht: D. Reidel.
  • Kuhn, T.S., 1970, The Structure of Scientific Revolutions , Chicago: University of Chicago Press, 2 nd edition.
  • –––, 1977, “Objectivity, Value Judgment, and Theory Choice,” in The Essential Tension: Selected Studies in Scientific Tradition and Change , T.S. Kuhn (ed.), Chicago: University of Chicago Press, pp. 320–339.
  • Ladyman, J., 2014, “Structural Realism”, The Stanford Encyclopedia of Philosophy (Spring 2014 Edition), E. N. Zalta (ed.), URL = < https://plato.stanford.edu/archives/spr2014/entries/structural-realism/ >.
  • Ladyman, J., O. Bueno, M. Suárez, and B. van Fraassen, 2011, “Scientific Representation: A Long Journey from Pragmatics to Pragmatics,” Metascience , 20 (3): 417–442.
  • Lakatos, I., 1980, The Methodology of Scientific Research Programmes (Philosophical Papers: Volume 1), Cambridge: Cambridge University Press.
  • Laudan, L., 1977, Progress and Its Problems: Towards a Theory of Scientific Growth , Berkeley, CA: University of California Press.
  • Leonelli, S., 2008, “Performing Abstraction: Two Ways of Modelling Arabidopsis thaliana ,” Biology and Philosophy , 23 (4): 509–528.
  • Levins, R., 1966, “The Strategy of Model Building in Population Biology,” American Scientist , 54 (4): 421–431.
  • Levins, R. and R. Lewontin, 1985, The Dialectical Biologist , Cambridge, MA: Harvard University Press.
  • Lewis, R.W., 1980, “Evolution: A System of Theories,” Perspectives in Biology and Medicine , 23 (4): 551–572.
  • Lewontin, R.C., 1972, “Apportionment of Human Diversity,” Evolutionary Biology , 6: 381–398.
  • –––, 1974, The Genetic Basis of Evolutionary Change , New York: Columbia University Press.
  • Lloyd, E., 1983, “The Nature of Darwin’s Support for the Theory of Natural Selection,” Philosophy of Science , 50 (1): 112–129.
  • –––, 1994 [1988], The Structure and Confirmation of Evolutionary Theory , Princeton: Princeton University Press.
  • –––, 2013 In Press, “Structure of Evolutionary Theory,” in International Encyclopedia of Social and Behavioral Sciences , W. Durham (ed.), 2 nd edition, Amsterdam: Elsevier.
  • London, F. and H. London, 1935, “The Electromagnetic Equations of the Supraconductor,” Proceedings of the Royal Society of London. Series A, Mathematical and Physical Sciences , 149 (866): 71–88.
  • Longino, H.E., 1995, “Gender, Politics, and the Theoretical Virtues,” Synthese 104 (3): 383–397.
  • –––, 2002, The Fate of Knowledge , Princeton: Princeton University Press.
  • –––, 2013, Studying Human Behavior: How Scientists Investigate Aggression & Sexuality , Chicago: University of Chicago Press.
  • López Beltrán, C., 1987, “La Explicación Evolucionista y el Uso de Modelos,” Masters Thesis, Posgrado en Filosofía de la Ciencia, Universidad Autónoma Metropolitana (Iztapalapa).
  • Lorenzano, P., 2013, “The Semantic Conception and the Structuralist View of Theories: A Critique of Suppe’s Criticisms,” Studies in History and Philosophy of Science (Part A) , 44: 600–607.
  • –––, 2014, “What is the Status of the Hardy-Weinberg Law within Population Genetics?,” in European Philosophy of Science: Philosophy of Science in Europe and the Viennese Heritage (Vienna Circle Institute Yearbook: Volume 17), M.C. Galavotti, E. Nemeth, F.L. Stadler F. (eds.), Cham, Switzerland: Springer, pp. 159–172.
  • Lowry, I., 1965, “A Short Course in Model Design,” Journal of the American Institute of Planners , 31 (2): 158–166.
  • Ludwig, D., 2015. “Against the New Metaphysics of Race,” Philosophy of Science 82: 1–21.
  • Lutz, S., 2012, “On a Straw Man in the Philosophy of Science: A Defense of the Received View,” HOPOS: The Journal of the International Society for the History of Philosophy of Science , 2 (1): 77–120.
  • –––, 2014, “What’s Right with a Syntactic Approach to Theories and Models?” Erkenntnis , 79 (8 supplement): 1475–1492.
  • –––, 2017, What “Was the Syntax-Semantics Debate in the Philosophy of Science About?,” Philosophy and Phenomenological Research , 95 (2): 319–352.
  • Mancosu, P., 2010, “Mathematical Style”, The Stanford Encyclopedia of Philosophy (Spring 2010 Edition), E. N. Zalta (ed.), URL = < https://plato.stanford.edu/archives/spr2010/entries/mathematical-style/ >.
  • Margenau, H., 1950, The Nature of Physical Reality: A Philosophy of Modern Physics , New York: McGraw-Hill.
  • Marker, D., 2002, Model Theory: An Introduction , New York: Springer.
  • Martínez, S., 2003, Geografía de las prácticas científicas: Racionalidad, heurística y normatividad , Mexico City: UNAM Press.
  • –––, 2014, “Technological Scaffolds for Culture and Cognition,” in Developing Scaffolds in Evolution, Culture and Cognition , L. Caporael, J. Griesemer, and W. Wimsatt (eds.), Cambridge, MA: MIT Press, pp. 249–264.
  • Matheson, C. and J. Dallmann, 2014, “Historicist Theories of Scientific Rationality”, The Stanford Encyclopedia of Philosophy (Fall 2014 Edition), E. N. Zalta (ed.), URL = < https://plato.stanford.edu/archives/fall2014/entries/rationality-historicist/ >.
  • McKinsey, J.C.C., A.C. Sugar, and P. Suppes, 1953, “Axiomatic Foundations of Classical Particle Mechanics,” Journal of Rational Mechanics and Analysis , 2 (2): 253–272.
  • Minsky, M., 1965, “Matter, Mind, and Models,” in Proceedings of the International Federation for Information Processing Congress (Volume 1), W. Kalenich (ed.), Washington D.C.: Spartan Books, pp. 45–49.
  • Morgan, M., 2012, The World in the Model: How Economists Work and Think , New York: Cambridge University Press.
  • Morgan, M.S. and M. Morrison (eds.), 1999, Models as Mediators: Perspectives on Natural and Social Science , Cambridge: Cambridge University Press.
  • Mormann, T., 2007, “The Structure of Scientific Theories in Logical Empiricism,” The Cambridge Companion to Logical Empiricism , in A. Richardson and T. Uebel (eds.), Cambridge: Cambridge University Press, pp. 136–162.
  • Morrison, M., 2007, “Where Have All the Theories Gone?,” Philosophy of Science , 74 (2): 195–228.
  • Moulines, C., 1976, “Approximate Application of Empirical Theories: A General Explication,” Erkenntnis , 10 (2): 201–227.
  • –––, 2002, “Introduction: Structuralism as a Program for Modelling Theoretical Science,” Synthese , 130 (1): 1–11.
  • Nagel, E., 1961, The Structure of Science: Problems in the Logic of Scientific Explanation , New York: Harcourt, Brace & World.
  • –––, 1979, “Issues in the Logic of Reductive Explanations,” in Teleology Revisited and Other Essays in the Philosophy and History of Science , New York: Columbia University Press, pp. 95–117.
  • Neurath, O., 1932, “Protokollsätze”, Erkenntnis , 3: 204–214; “Protocol Statements,” in Philosophical Papers 1913-1946 , R.S. Cohen and M. Neurath (eds.), Dordrecht: Reidel (1983), pp. 91–99.
  • Nicholson, D. and R. Gawne, 2014, “Rethinking Woodger’s Legacy in the Philosophy of Biology,” Journal of the History of Biology , 47 (2): 243–292.
  • Nolte, D.D., 2010, “The Tangled Tale of Phase Space,” Physics Today , April: 33–38.
  • Okasha, S., 2012, “Population Genetics”, The Stanford Encyclopedia of Philosophy (Fall 2012 Edition), E. N. Zalta (ed.), URL = < https://plato.stanford.edu/archives/fall2012/entries/population-genetics/ >.
  • Oppenheimer, J.R., 1956, “Analogy in Science,” American Psychologist , 11 (3): 127–135.
  • Oyama, S., 2000, The Ontogeny of Information: Developmental Systems and Evolution , 2 nd ed., Durham: Duke University Press.
  • Pereda, C., 2013, “Ulises Moulines y la concepción estructural de las teorías científicas,” in La filosofía en México en el siglo XX: Apuntes de un participante , C. Pereda, Mexico City: CONACULTA (Consejo Nacional para la Cultura y las Artes), pp. 200–212.
  • Pickstone, J.V., 2000, Ways of Knowing: A New History of Science, Technology and Medicine , Chicago: University of Chicago Press.
  • Pigliucci, M. and G.B. Müller, 2010, Evolution: The Extended Synthesis , Cambridge, MA: MIT Press.
  • Popper, K., 1996 [1976], “The Myth of the Framework,” In The Myth of the Framework: In Defence of Science and Rationality , M. A. Notturno (ed), Abingdon: Routledge, pp. 33–64.
  • Pritchard J.K., M. Stephens, and P. Donnelly, 2000, “Inference of Population Structure Using Multilocus Genotype Data,” Genetics , 155 (2): 945–959.
  • Preston, J., 2012, “Paul Feyerabend”, The Stanford Encyclopedia of Philosophy (Winter 2012 Edition), E. N. Zalta (ed.), URL = < https://plato.stanford.edu/archives/win2012/entries/feyerabend/ >.
  • Przełęcki, M., 1969, The Logic of Empirical Theories , London: Routledge & Kegan Paul.
  • Putnam, H., 1962, “What Theories Are Not,” in Logic, Methodology, and Philosophy of Science: Proceedings of the 1960 International Congress , E. Nagel, P. Suppes, and A. Tarski (eds.), Stanford, CA: Stanford University Press, pp. 240–251.
  • Reichenbach, H., 1938, Experience and Prediction: An Analysis of the Foundations and the Structure of Knowledge , Chicago: University of Chicago Press.
  • –––, 1965 [1920], The Theory of Relativity and A Priori Knowledge , with an introduction by M. Reichenbach, Berkeley: University of California Press. Original: Relativitätstheorie und Erkenntnis apriori , Berlin: Springer.
  • –––, 1969 [1924], The Axiomatization of the Theory of Relativity , with an introduction by W.C. Salmon. Berkeley-Los Angeles: University of California Press. Original: Axiomatik der relativistischen Raum-Zeit-Lehre , Braunschweig: F. Vieweg & Sohn.
  • –––, 1978, Selected Writings, 1909–1953: With a Selection of Biographical and Autobiographical Sketches (Volumes 1–2), Dordrecht: Reidel.
  • Rice, S., 2004, Evolutionary Theory: Mathematical and Conceptual Foundations , Sunderland, MA: Sinauer Associates.
  • Richards, R., 1992, “The Structure of Narrative Explanation in History and Biology,” in History and Evolution , M. Nitecki and D. Nitecki (eds.), Albany: State University of New York Press, pp. 19–53.
  • Richardson, A., 2002, “Engineering Philosophy of Science: American Pragmatism and Logical Empiricism in the 1930s,” Philosophy of Science , 69 (S3): S36–S47.
  • Rosenberg N.A., J.K. Pritchard, J.L. Weber, H.M. Cann, K.K. Kidd, L.A. Zhivotovsky, and M.A. Feldman, 2002, “Genetic Structure of Human Populations,” Science , 298 (5602): 2381–2385.
  • Rosenblueth, A. and N. Wiener, 1945, “The Role of Models in Science,” Philosophy of Science , 12 (4): 316–321.
  • Ruse, M., 1975, “Charles Darwin’s Theory of Evolution: An Analysis,” Journal of the History of Biology , 8 (2): 219–241.
  • Rutte, H., 1991, “Neurath contra Schlick. On the Discussion of Truth in the Vienna Circle,” in Rediscovering the Forgotten Vienna Circle: Austrian studies on Otto Neurath and the Vienna Circle , T. Uebel (ed.), Dordrecht: Kluwer, pp. 169–174.
  • Sarkar, S., 1998, Genetics and Reductionism , Cambridge: Cambridge University Press.
  • Savage, C.W., 1990, “Preface,” in Scientific Theories. Minnesota Studies in the Philosophy of Science. Volume 14, C.W. Savage (ed.), Minneapolis: University of Minnesota Press, pp. vii–ix.
  • Schaffner K., 1969, “Correspondence Rules,” Philosophy of Science , 36 (3): 280–290.
  • –––, 1976, “Reductionism in Biology: Prospects and Problems,” in PSA : Proceedings of the Biennial Meeting of the Philosophy of Science Association 1974 : 613–632.
  • –––, 1993, Discovery and Explanation in Biology and Medicine , Chicago: University of Chicago Press.
  • Schlick, M., 1925 [1918], General Theory of Knowledge , LaSalle, IL: Open Court.
  • –––, 1934, “Über das Fundament der Erkenntnis,” Erkenntnis , 4 (1): 79–99.
  • Schmidt, H.-J., 2014, “Structuralism in Physics”, The Stanford Encyclopedia of Philosophy (Spring 2014 Edition), E. N. Zalta (ed.), URL = < https://plato.stanford.edu/archives/spr2014/entries/physics-structuralism/ >.
  • Shapin, S. and S. Schaffer, 1985, Leviathan and the Air-Pump: Hobbes, Boyle, and the Experimental Life , Princeton: Princeton University Press.
  • Simon, H., 1954, “The Axiomatization of Classical Mechanics,” Philosophy of Science , 21 (4): 340–343.
  • –––, 1957, Models of Man , New York: Wiley.
  • –––, 1970, “The Axiomatization of Physical Theories,” Philosophy of Science , 37 (1): 16–26.
  • Smith, B.C., 1996, On the Origin of Objects , Cambridge, MA: MIT Press.
  • Sneed, J., 1979, The Logical Structure of Mathematical Physics , Dordrecht: D. Reidel, 2 nd edition.
  • Spencer, Q., 2015, “Philosophy of Race Meets Population Genetics,” Studies in History and Philosophy of Biological and Biomedical Sciences 52: 46–55.
  • Stegmüller, W., 1976, The Structure and Dynamics of Theories , New York: Springer.
  • –––, 1979, “The Structuralist View: Survey, Recent Developments and Answers to Some Criticisms”, in The Logic and Epistemology of Scientific Change , I. Niiniluoto and R. Tuomela (eds.), Amsterdam: North Holland.
  • Suárez, M., 1999, “The Role of Models in the Application of Scientific Theories; Epistemological Implications,” in Models as Mediators. Perspectives on Natural and Social Science , M.S. Morgan and M. Morrison (eds.), Cambridge: Cambridge University Press, pp. 168–196.
  • –––, 2011, Comment on van Fraassen Scientific Representation: Paradoxes of Perspective , in Ladyman, J., O. Bueno, M. Suárez, and B. van Fraassen, “Scientific Representation: A Long Journey from Pragmatics to Pragmatics,” Metascience , 20 (3): 428–433.
  • Suárez, M. and N. Cartwright, 2008, “Theories: Tools versus Models,” Studies in History and Philosophy of Modern Physics , 39 (1): 62–81.
  • Suárez, M. and F. Pero, 2019, “The Representational Semantic Conception,” Philosophy of Science , 86 (2): 344–365.
  • Suppe, F., 1977, The Structure of Scientific Theories , Urbana, IL: University of Illinois Press.
  • –––, 1989, The Semantic Conception of Theories and Scientific Realism , Chicago: University of Illinois Press.
  • –––, 2000, “Understanding Scientific Theories: An Assessment of Developments,” PSA: Proceedings of the Biennial Meeting of the Philosophy of Science Association 1998 , (2): S102–S115.
  • Suppes, P., 1957, Introduction to Logic , Princeton: D. Van Nostrand Co.
  • –––, 1960, “A Comparison of the Meaning and Uses of Models in Mathematics and the Empirical Sciences,” Synthese , 12 (2-3): 287–301.
  • –––, 1962, “Models of Data,” in Logic, Methodology, and Philosophy of Science: Proceedings of the 1960 International Congress , E. Nagel, P. Suppes, and A. Tarski (eds.), Stanford, CA: Stanford University Press, pp. 252–261.
  • –––, 1967, “What is a Scientific Theory?,” In Philosophy of Science Today , S. Morgenbesser (ed.), New York: Basic Books, pp. 55–67.
  • –––, 1968, “The Desirability of Formalization in Science,” The Journal of Philosophy , 65 (20): 651–664.
  • –––, 1978, “The Plurality of Science,” PSA: Proceedings of the Biennial Meeting of the Philosophy of Science Association 1978 , (2): 3–16.
  • –––, 2002, Representation and Invariance of Scientific Structures , Stanford, CA: CSLI Publications.
  • Swoyer, C., 2014, “Relativism”, The Stanford Encyclopedia of Philosophy (Winter 2014 Edition), E. N. Zalta (ed.), URL = < https://plato.stanford.edu/archives/win2014/entries/relativism/ >.
  • Thompson, P., 1989, The Structure of Biological Theories , Albany: SUNY Press.
  • –––, 2007, “Formalisations of Evolutionary Biology,” in Philosophy of Biology , M. Matthen and C. Stephens (eds.), Elsevier, Amsterdam, pp. 485–523
  • Thomson-Jones, M., 2012, “Modelling without Mathematics,” Philosophy of Science , 79 (5): 761–772. (Same individual as Jones 2005.)
  • Toulmin, S., 1972, Human Understanding: The Collective Use and Evolution of Concepts , Princeton: Princeton University Press.
  • Tuomi, J., 1981, “Structure and Dynamics of Darwinian Evolutionary Theory,” Systematic Zoology , 30 (1): 22–31.
  • –––, 1992, “Evolutionary Synthesis: A Search for the Strategy,” Philosophy of Science , 59 (3): 429–438.
  • Tversky, A., 1977, “Features of Similarity,” Psychological Review , 84 (4): 327–352.
  • Uebel, T., 2014, “Vienna Circle”, The Stanford Encyclopedia of Philosophy (Spring 2014 Edition), E. N. Zalta (ed.), URL = < https://plato.stanford.edu/archives/spr2014/entries/vienna-circle/ >.
  • van Benthem J., 2012, “The Logic of Empirical Theories Revisited,” Synthese , 186 (3): 775–792.
  • van Fraassen, B., 1967, “Meaning Relations among Predicates,” Noûs , 1 (2): 161–179.
  • –––, 1970, “On the Extension of Beth’s Semantics of Physical Theories,” Philosophy of Science , 37 (3): 325–339.
  • –––, 1980, The Scientific Image , Oxford: Oxford University Press.
  • –––, 1981, “Theory Construction and Experiment: An Empiricist View,” PSA: Proceedings of the Biennial Meeting of the Philosophy of Science Association 1980 , (2): 663–678.
  • –––, 1989, Laws and Symmetry , New York: Oxford University Press.
  • –––, 2008, Scientific Representation: Paradoxes of Perspective , New York: Oxford University Press.
  • van Riel, R. and R. Van Gulick, 2014, “Scientific Reduction”, The Stanford Encyclopedia of Philosophy (Summer 2014 Edition), E. N. Zalta (ed.), URL = < https://plato.stanford.edu/archives/sum2014/entries/scientific-reduction/ >.
  • Van Valen, L., 1976, “Domains, Deduction, the Predictive Method, and Darwin,” Evolutionary Theory , 1: 231–245.
  • Vicedo, M., 1995, “Scientific Styles: Toward Some Common Ground in the History, Philosophy, and Sociology of Science,” Perspectives on Science , 3: 231–254.
  • Vickers, P., 2009, “Can Partial Structures Accommodate Inconsistent Science?” Principia , 13 (2): 233–250.
  • Walsh, D., 2015, Organisms, Agency, and Evolution, Cambridge: Cambridge University Press.
  • Weisberg, M., 2013, Simulation and Similarity: Using Models to Understand the World , New York: Oxford University Press.
  • Wessels, L., 1976, “Laws and Meaning Postulates in van Fraassen’s View of Theories,” in PSA: Proceedings of the Biennial Meeting of the Philosophy of Science Association 1974 : 215–234.
  • Williams, M., 1970, “Deducing the Consequences of Selection: A Mathematical Model,” Journal of Theoretical Biology , 48: 343–385.
  • –––, 1973, “The Logical Status of Natural Selection and other Evolutionary Controversies: Resolution by Axiomatization,” in M. Bunge (ed.), The Methodological Unity of Science , Dordrecht: D. Reidel, pp. 84–102.
  • Wimsatt, W.C., 2007, Re-Engineering Philosophy for Limited Beings: Piecewise Approximations to Reality , Cambridge, MA: Harvard University Press.
  • Winsberg, E., 2010, Science in the Age of Computer Simulation , Chicago: University of Chicago Press.
  • –––, 2018, Philosophy and Climate Science , Cambridge: Cambridge University Press.
  • Winther, R.G., 2006a, “Parts and Theories in Compositional Biology,” Biology and Philosophy , 21 (4): 471–499.
  • –––, 2006b, “Fisherian and Wrightian Perspectives in Evolutionary Genetics and Model-Mediated Imposition of Theoretical Assumptions,” Journal of Theoretical Biology , 240 (2): 218–232.
  • –––, 2009, “Schaffner’s Model of Theory Reduction: Critique and Reconstruction,” Philosophy of Science , 76 (2): 119–142.
  • –––, 2011, “Part-Whole Science,” Synthese , 178 (3): 397–427.
  • –––, 2012a, “Mathematical Modeling in Biology: Philosophy and Pragmatics,” Frontiers in Plant Evolution and Development , 3: 102, doi:10.3389/fpls.2012.00102
  • –––, 2012b, “Interweaving Categories: Styles, Paradigms, and Models,” Studies in History and Philosophy of Science (Part A) , 43 (4): 628–639.
  • –––, 2014, “The Genetic Reification of ‘Race’? A Story of Two Mathematical Methods,” Critical Philosophy of Race , 2 (2): 204–223.
  • –––, 2020, When Maps Become the World , Chicago, IL: University of Chicago Press.
  • Winther, R.G., R. Giordano, M.D. Edge, and R. Nielsen, 2015, “The Mind, the Lab, and the Field: Three Kinds of Populations in Scientific Practice,” Studies in History and Philosophy of Biological and Biomedical Sciences , 52: 12–21.
  • Winther, R.G. and J.M. Kaplan, 2013, “Ontologies and Politics of Biogenomic ‘Race’,” Theoria. A Journal of Social and Political Theory (South Africa) , 60 (3): 54–80.
  • Woodger J.H., 1937, The Axiomatic Method in Biology , Cambridge: Cambridge University Press.
  • –––, 1959, “Studies in the Foundations of Genetics,” in The Axiomatic Method with Special Reference to Geometry and Physics: Proceedings of an International Symposium Held at the University of California, Berkeley, December 26, 1957 – January 4, 1958 , L. Henkin, P. Suppes, and A. Tarski (eds.), Amsterdam: North Holland, pp. 408–428.
  • Worrall, J., 1984, “An Unreal Image,” The British Journal for the Philosophy of Science , 35 (1): 65–80.
  • Wright, S., 1969, Evolution and the Genetics of Populations: A Treatise in Four Volumes (Volume 2: The Theory of Gene Frequencies), Chicago: University of Chicago Press.
  • Zach, R., 2009, “Hilbert’s Program”, The Stanford Encyclopedia of Philosophy (Spring 2009 Edition), E. N. Zalta (ed.), URL = < https://plato.stanford.edu/archives/spr2009/entries/hilbert-program/ >.
  • Ziman, J., 2000, Real Science: What It Is, and What It Means , Cambridge: Cambridge University Press.
How to cite this entry . Preview the PDF version of this entry at the Friends of the SEP Society . Look up topics and thinkers related to this entry at the Internet Philosophy Ontology Project (InPhO). Enhanced bibliography for this entry at PhilPapers , with links to its database.
  • Koellner, P., ms., “ Carnap on the Foundations of Logic and Mathematics ,” unpublished.
  • Browse Philpapers on The Nature of Theories
  • Browse Philpapers on Theoretical Virtues
  • Browser Philpapers on Models and Idealization
  • Evolution Resources from the National Academies
  • Definitions of Fact, Theory, and Law in Scientific Work , National Center for Science Education (NCSE).

Carnap, Rudolf | cognitive science | confirmation | Darwinism | empiricism: logical | feminist philosophy, interventions: epistemology and philosophy of science | Feyerabend, Paul | genetics: population | incommensurability: of scientific theories | Kuhn, Thomas | models in science | model theory | paradox: Skolem’s | physics: structuralism in | pragmatism | rationality: historicist theories of | reduction, scientific | science: theory and observation in | scientific explanation | scientific realism | scientific representation | simulations in science | statistical physics: philosophy of statistical mechanics | structural realism | style: in mathematics | theoretical terms in science | underdetermination, of scientific theories | Vienna Circle

Acknowledgments

The following provided helpful feedback or conversation, or both, Jácome Armas, Nancy Cartwright, Mario Casanueva, Carl Craver, Eugene Earnshaw, Doc Edge, Michael Friedman, Sari Friedman, Fermín Fulda, Ryan Giordano, Ian Hacking, Hervé Kieffel, Elisabeth A. Lloyd, Helen Longino, Carlos López Beltrán, Greg Lusk, Sebastian Lutz, Sergio Martínez, Amir Najmi, Thomas Ryckman, Mette Bannergaard Johansen, Mette Smølz Skau, Bas van Fraassen, Denis Walsh, Ole Wæver, and two anonymous reviewers. Alex Dor, Cory Knudson, and Lucas McGranahan offered expert research assistance.

Copyright © 2020 by Rasmus Grønfeldt Winther < rgwinther @ gmail . com >

  • Accessibility

Support SEP

Mirror sites.

View this site from another server:

  • Info about mirror sites

The Stanford Encyclopedia of Philosophy is copyright © 2023 by The Metaphysics Research Lab , Department of Philosophy, Stanford University

Library of Congress Catalog Data: ISSN 1095-5054

Hypothesis, Model, Theory, and Law

Dorling Kindersley / Getty Images

  • Physics Laws, Concepts, and Principles
  • Quantum Physics
  • Important Physicists
  • Thermodynamics
  • Cosmology & Astrophysics
  • Weather & Climate

hypothesis and scientific theory

  • M.S., Mathematics Education, Indiana University
  • B.A., Physics, Wabash College

In common usage, the words hypothesis, model, theory, and law have different interpretations and are at times used without precision, but in science they have very exact meanings.

Perhaps the most difficult and intriguing step is the development of a specific, testable hypothesis. A useful hypothesis enables predictions by applying deductive reasoning, often in the form of mathematical analysis. It is a limited statement regarding the cause and effect in a specific situation, which can be tested by experimentation and observation or by statistical analysis of the probabilities from the data obtained. The outcome of the test hypothesis should be currently unknown, so that the results can provide useful data regarding the validity of the hypothesis.

Sometimes a hypothesis is developed that must wait for new knowledge or technology to be testable. The concept of atoms was proposed by the ancient Greeks , who had no means of testing it. Centuries later, when more knowledge became available, the hypothesis gained support and was eventually accepted by the scientific community, though it has had to be amended many times over the year. Atoms are not indivisible, as the Greeks supposed.

A model is used for situations when it is known that the hypothesis has a limitation on its validity. The Bohr model of the atom , for example, depicts electrons circling the atomic nucleus in a fashion similar to planets in the solar system. This model is useful in determining the energies of the quantum states of the electron in the simple hydrogen atom, but it is by no means represents the true nature of the atom. Scientists (and science students) often use such idealized models  to get an initial grasp on analyzing complex situations.

Theory and Law

A scientific theory or law represents a hypothesis (or group of related hypotheses) which has been confirmed through repeated testing, almost always conducted over a span of many years. Generally, a theory is an explanation for a set of related phenomena, like the theory of evolution or the big bang theory . 

The word "law" is often invoked in reference to a specific mathematical equation that relates the different elements within a theory. Pascal's Law refers an equation that describes differences in pressure based on height. In the overall theory of universal gravitation developed by Sir Isaac Newton , the key equation that describes the gravitational attraction between two objects is called the law of gravity .

These days, physicists rarely apply the word "law" to their ideas. In part, this is because so many of the previous "laws of nature" were found to be not so much laws as guidelines, that work well within certain parameters but not within others.

Scientific Paradigms

Once a scientific theory is established, it is very hard to get the scientific community to discard it. In physics, the concept of ether as a medium for light wave transmission ran into serious opposition in the late 1800s, but it was not disregarded until the early 1900s, when Albert Einstein proposed alternate explanations for the wave nature of light that did not rely upon a medium for transmission.

The science philosopher Thomas Kuhn developed the term scientific paradigm to explain the working set of theories under which science operates. He did extensive work on the scientific revolutions that take place when one paradigm is overturned in favor of a new set of theories. His work suggests that the very nature of science changes when these paradigms are significantly different. The nature of physics prior to relativity and quantum mechanics is fundamentally different from that after their discovery, just as biology prior to Darwin’s Theory of Evolution is fundamentally different from the biology that followed it. The very nature of the inquiry changes.

One consequence of the scientific method is to try to maintain consistency in the inquiry when these revolutions occur and to avoid attempts to overthrow existing paradigms on ideological grounds.

Occam’s Razor

One principle of note in regards to the scientific method is Occam’s Razor (alternately spelled Ockham's Razor), which is named after the 14th century English logician and Franciscan friar William of Ockham. Occam did not create the concept—the work of Thomas Aquinas and even Aristotle referred to some form of it. The name was first attributed to him (to our knowledge) in the 1800s, indicating that he must have espoused the philosophy enough that his name became associated with it.

The Razor is often stated in Latin as:

entia non sunt multiplicanda praeter necessitatem
or, translated to English:
entities should not be multiplied beyond necessity

Occam's Razor indicates that the most simple explanation that fits the available data is the one which is preferable. Assuming that two hypotheses presented have equal predictive power, the one which makes the fewest assumptions and hypothetical entities takes precedence. This appeal to simplicity has been adopted by most of science, and is invoked in this popular quote by Albert Einstein:

Everything should be made as simple as possible, but not simpler.

It is significant to note that Occam's Razor does not prove that the simpler hypothesis is, indeed, the true explanation of how nature behaves. Scientific principles should be as simple as possible, but that's no proof that nature itself is simple.

However, it is generally the case that when a more complex system is at work there is some element of the evidence which doesn't fit the simpler hypothesis, so Occam's Razor is rarely wrong as it deals only with hypotheses of purely equal predictive power. The predictive power is more important than the simplicity.

Edited by Anne Marie Helmenstine, Ph.D.

  • Scientific Hypothesis, Model, Theory, and Law
  • Einstein's Theory of Relativity
  • What Is a Paradigm Shift?
  • Wave Particle Duality and How It Works
  • Oversimplification and Exaggeration Fallacies
  • Kinetic Molecular Theory of Gases
  • Scientific Method
  • Understanding Cosmology and Its Impact
  • The History of Gravity
  • The Copenhagen Interpretation of Quantum Mechanics
  • Tips on Winning the Debate on Evolution
  • Geological Thinking: Method of Multiple Working Hypotheses
  • The Life and Work of Albert Einstein
  • Newton's Law of Gravity
  • De Broglie Hypothesis
  • What Is a Scientific or Natural Law?

What is a scientific theory?

A scientific theory is based on careful examination of facts.

scientific theory: a chalkboard being drawn on

  • The process
  • Good theory characteristics

The difference between theories, facts and laws

Additional resources, bibliography.

A scientific theory is a structured explanation to explain a group of facts or phenomena in the natural world that often incorporates a scientific hypothesis and scientific laws . The scientific definition of a theory contrasts with the definition most people use in casual language.

"The way that scientists use the word 'theory' is a little different than how it is commonly used in the lay public," said Jaime Tanner, a professor of biology at Emerson College in Boston. "Most people use the word 'theory' to mean an idea or hunch that someone has, but in science the word 'theory' refers to the way that we interpret facts."

Related: 5 sci-fi concepts that are possible (in theory)

The process of becoming a scientific theory

Every scientific theory relies on the scientific method . A scientist may make an observation and devise a hypothesis to explain that observation, then design an experiment to test that hypothesis. If the hypothesis is shown to be incorrect, the scientist will develop a new hypothesis and begin the process again. If the hypothesis is supported by the results of the experiment, it will go on to be tested again. If the hypothesis isn't disproven or surpassed by a better explanation, the scientist may incorporate it into a larger theory that helps to explain the observed phenomenon and relates it to other phenomena, according to the Field Museum . 

A scientific theory is not the end result of the scientific method; theories can be proven or rejected, just like hypotheses . And theories are continually improved or modified as more information is gathered, so that the accuracy of the prediction becomes greater over time.

Theories are foundations for furthering scientific knowledge and for putting the information gathered to practical use. Scientists use theories to develop inventions or find a cure for a disease.

Furthermore, a scientific theory is the framework for observations and facts, Tanner said. Theories may change, or the way that they are interpreted may change, but the facts themselves don't change. Tanner likens theories to a basket in which scientists keep facts and observations that they find. The shape of that basket may change as the scientists learn more and include more facts. "For example, we have ample evidence of traits in populations becoming more or less common over time (evolution), so evolution is a fact, but the overarching theories about evolution, the way that we think all of the facts go together might change as new observations of evolution are made," Tanner told Live Science.

Characteristics of a good theory

The University of California, Berkeley , defines a theory as "a broad, natural explanation for a wide range of phenomena. Theories are concise, coherent, systematic, predictive, and broadly applicable, often integrating and generalizing many hypotheses." 

According to Columbia University emeritus professor of philosophy Philip Kitcher, a good scientific theory has three characteristics. First, it has unity, which means it consists of a limited number of problem-solving strategies that can be applied to a wide range of scientific circumstances. Second, a good scientific theory leads to new questions and new areas of research. This means that a theory doesn't need to explain everything in order to be useful. And finally, a good theory is formed from a number of hypotheses that can be tested independently from the theory itself.

Any scientific theory must be based on a careful and rational examination of the facts. Facts and theories are two different things. In the scientific method, there is a clear distinction between facts, which can be observed and/or measured, and theories, which are scientists' explanations and interpretations of the facts. 

Some think that theories become laws, but theories and laws have separate and distinct roles in the scientific method. A law is a description of an observed phenomenon in the natural world that holds true every time it is tested. It doesn't explain why something is true; it just states that it is true. A theory, on the other hand, explains observations that are gathered during the scientific process. So, while law and theory are part of the scientific process, they are two different aspects, according to the National Center for Science Education . 

A good example of the difference between a theory and a law is the case of Gregor Mendel . In his research, Mendel discovered that two separate genetic traits would appear independently of each other in different offspring. "Yet, Mendel knew nothing of DNA or chromosomes . It wasn't until a century later that scientists discovered DNA and chromosomes — the biochemical explanation of Mendel's laws," said Peter Coppinger, an associate professor of biology and biomedical engineering at the Rose-Hulman Institute of Technology. "It was only then that scientists, such as T.H. Morgan working with fruit flies, explained the Law of Independent Assortment using the theory of chromosomal inheritance. Still today, this is the universally accepted explanation [theory] for Mendel's Law."

  • When does a theory become a fact? This article from Arizona State University says you're asking the wrong question! 
  • Learn the difference between the casual and scientific uses of "theory" and "law" from the cartoony stars of the Amoeba Sisters on Youtube.
  • Can a scientific theory be falsified? This article from Scientific American says no. 

Kenneth Angielczyk, "What Do We Mean by "Theory" in Science?" Field Museum, March 10, 2017. https://www.fieldmuseum.org/blog/what-do-we-mean-theory-science

University of California, Berkeley, "Science at multiple levels." https://undsci.berkeley.edu/article/0_0_0/howscienceworks_19  

Philip Kitcher, "Abusing Science: The Case Against Creationism," MIT Press, 1982. 

National Center for Science Education, "Definitions of Fact, Theory, and Law in Scientific Work," March 16, 2016 https://ncse.ngo/definitions-fact-theory-and-law-scientific-work  

Sign up for the Live Science daily newsletter now

Get the world’s most fascinating discoveries delivered straight to your inbox.

Alina Bradford

Why do babies rub their eyes when they're tired?

Why do people dissociate during traumatic events?

Longest eclipse ever: How scientists rode the supersonic Concorde jet to see a 74-minute totality

Most Popular

By Anna Gora December 27, 2023

By Anna Gora December 26, 2023

By Anna Gora December 25, 2023

By Emily Cooke December 23, 2023

By Victoria Atkinson December 22, 2023

By Anna Gora December 16, 2023

By Anna Gora December 15, 2023

By Anna Gora November 09, 2023

By Donavyn Coffey November 06, 2023

By Anna Gora October 31, 2023

By Anna Gora October 26, 2023

  • 2 Error-corrected qubits 800 times more reliable after breakthrough, paving the way for 'next level' of quantum computing
  • 3 April 8 solar eclipse: What time does totality start in every state?
  • 4 Giant coyote killed in southern Michigan turns out to be a gray wolf — despite the species vanishing from region 100 years ago
  • 5 James Webb telescope confirms there is something seriously wrong with our understanding of the universe
  • 2 Giant coyote killed in southern Michigan turns out to be a gray wolf — despite the species vanishing from region 100 years ago
  • 3 James Webb telescope confirms there is something seriously wrong with our understanding of the universe
  • 4 Cholesterol-gobbling gut bacteria could protect against heart disease
  • 5 'Gambling with your life': Experts weigh in on dangers of the Wim Hof method

hypothesis and scientific theory

Library homepage

  • school Campus Bookshelves
  • menu_book Bookshelves
  • perm_media Learning Objects
  • login Login
  • how_to_reg Request Instructor Account
  • hub Instructor Commons
  • Download Page (PDF)
  • Download Full Book (PDF)
  • Periodic Table
  • Physics Constants
  • Scientific Calculator
  • Reference & Cite
  • Tools expand_more
  • Readability

selected template will load here

This action is not available.

Physics LibreTexts

1.2: Theories, Hypotheses and Models

  • Last updated
  • Save as PDF
  • Page ID 19359

For the purpose of this textbook (and science in general), we introduce a distinction in what we mean by “theory”, “hypothesis”, and by “model”. We will consider a “theory” to be a set of statements (or an equation) that gives us a broad description, applicable to several phenomena and that allows us to make verifiable predictions. For example, Chloë’s Theory ( \(t \propto \sqrt{h}\) ) can be considered a theory. Specifically, we do not use the word theory in the context of “I have a theory about this...”

A “hypothesis” is a consequence of the theory that one can test. From Chloë’s Theory, we have the hypothesis that an object will take \(\sqrt{2}\) times longer to fall from \(1\:\text{m}\) than from \(2\:\text{m}\) . We can formulate the hypothesis based on the theory and then test that hypothesis. If the hypothesis is found to be invalidated by experiment, then either the theory is incorrect, or the hypothesis is not consistent with the theory.

A “model” is a situation-specific description of a phenomenon based on a theory , that allows us to make a specific prediction. Using the example from the previous section, our theory would be that the fall time of an object is proportional to the square root of the drop height, and a model would be applying that theory to describe a tennis ball falling by \(4.2\) m. From the model, we can form a testable hypothesis of how long it will take the tennis ball to fall that distance. It is important to note that a model will almost always be an approximation of the theory applied to describe a particular phenomenon. For example, if Chloë’s Theory is only valid in vacuum, and we use it to model the time that it take for an object to fall at the surface of the Earth, we may find that our model disagrees with experiment. We would not necessarily conclude that the theory is invalidated, if our model did not adequately apply the theory to describe the phenomenon (e.g. by forgetting to include the effect of air drag).

This textbook will introduce the theories from Classical Physics, which were mostly established and tested between the seventeenth and nineteenth centuries. We will take it as given that readers of this textbook are not likely to perform experiments that challenge those well-established theories. The main challenge will be, given a theory, to define a model that describes a particular situation, and then to test that model. This introductory physics course is thus focused on thinking of “doing physics” as the task of correctly modeling a situation.

Emma's Thoughts

What’s the difference between a model and a theory?

“Model” and “Theory” are sometimes used interchangeably among scientists. In physics, it is particularly important to distinguish between these two terms. A model provides an immediate understanding of something based on a theory.

For example, if you would like to model the launch of your toy rocket into space, you might run a computer simulation of the launch based on various theories of propulsion that you have learned. In this case, the model is the computer simulation, which describes what will happen to the rocket. This model depends on various theories that have been extensively tested such as Newton’s Laws of motion, Fluid dynamics, etc.

  • “Model”: Your homemade rocket computer simulation
  • “Theory”: Newton’s Laws of motion, Fluid dynamics

With this analogy, we can quickly see that the “model” and “theory” are not interchangeable. If they were, we would be saying that all of Newton’s Laws of Motion depend on the success of your piddly toy rocket computer simulation!

Exercise \(\PageIndex{2}\)

Models cannot be scientifically tested, only theories can be tested.

All The Differences

Hypothesis VS A Scientific Theory: A Comparison

Categories Philosophy

Hypothesis VS A Scientific Theory: A Comparison

A hypothesis is an explanation that is proposed for a phenomenon. If a hypothesis has to be scientific, the scientific way needs that it can be tested. Scientists basically base their hypotheses on prior observations that the available scientific theories can’t explain .

A scientific theory on the other hand, is an explanation of an aspect of the natural world or the universe which is tested and corroborated according to scientific ways. In situations where theories aren’t amenable to experimental testing, those theories will be evaluated through the principles of abductive reasoning.

One must keep in mind that even though ‘hypothesis’ and ‘theory’ are usually used interchangeably, both aren’t the same.

The difference between a hypothesis and a scientific theory is that a hypothesis is an assumption that is made prior to any research which has to be done. The hypothesis is tested in order to see if it is true or not. A theory is a principle that has to be formed in order to explain the things that are shown in data.

Here is a table for the differences between hypothesis and scientific theory:

Keep reading to know more.

Page Contents

What is a scientific theory?

lab samples being held by a person

A scientific theory explains an aspect of the natural world or the universe, the explanation is tested and corroborated according to the methods of science. In everyday speech, theory refers to an explanation that is a representation of an unsubstantiated and speculative guess. However, in science, the theory describes an explanation that is tested and is accepted as valid.

A scientific theory is different from a scientific fact and scientific law. A fact is merely a basic observation, while a law is a statement. Stephen Jay Gould  wrote that “Facts are the world’s data, while theories are structures of ideas which explain and interpret the facts.”

Learn the differences between fact, theory, hypothesis, and laws through this video.

A scientific theory can be rejected if it’s unfit for the new findings. Many well-established theories are unlikely to be ever changed, such as evolution and cell theory.

Scientific theories are very much testable and can form falsifiable predictions. Theories explain the cause of a certain natural phenomenon, they are also used to describe as well as predict particular aspects of the universe. Like other types of scientific knowledge, a scientific theory is deductive as well as inductive, which aims for predictive and explanatory power. Moreover, Scientists often utilize theories for further scientific information, and also to provide advancements to technology or medicine.

Furthermore, two types of scientific theories were described by Albert Einstein: “ Constructive theories” and “ Principle theories”. Constructive theories are the constructive representation of phenomena, such as kinetic theory. Principle theories, on the other hand, are empirical generalizations, for example, Newton’s laws of motion.

What is an example of a theory in science?

the earth as seen from space

The world is full of theories, and some unbelievable theories are true as a theory is formed to explain the things that exist.

Here are some popular examples of theories in science:

  • Big Bang Theory
  • Theory of Evolution
  • Quantum Field Theory
  • Hubble’s Law of Cosmic Expansion
  • Universal Law of Gravitation
  • Newton’s Laws of Motion
  • Laws of Thermodynamics
  • Archimedes’ Buoyancy Principle

What does a hypothesis mean?

a statue of Einstein

A hypothesis is often tentative, an assumption or suggestion that is formed strictly for the sole objective of being tested. A hypothesis is an explanation for a phenomenon. A hypothesis is based on prior observations that couldn’t be explained with the scientific theories already available.

The meaning of the adjective “hypothetical” is “Possessing the nature of a hypothesis” or “Assuming to exist as an immediate result of a hypothesis”.

People often refer to a hypothesis as an “educated guess” as it provides an outcome that is based on the evidence. Although, there are several scientists who reject the term “educated guess”. Moreover, experimenters have the right to test and reject the hypothesis prior to solving the problem.

Researchers, while weighing up alternative hypotheses, will take into consideration:

  • Testability.
  • Scope: The evident application of the hypothesis to several cases of phenomena.
  • Fruitfulness: the likelihood of a hypothesis to explain phenomena.
  • Conservatism: It’s the degree of “fit” with the existing known knowledge systems.

How do I write a hypothesis?

two scientists looking through a microscope

There are 6 steps that one has to follow to write a plausible hypothesis. Here is all the information that you would need.

  • Ask a question

Curiosity has been the inspiration for many of history’s greatest scientific achievements, so the best place to begin is to ask yourself questions about the world as well as the universe. Why are things the way they are? You must choose a research topic that is interesting to you, and your curiosity will come naturally. 

  • Conduct preliminary research

Gather all the background information on your research topic. How much background information you require depends on what you are attempting and what topic you chose. You might need to read several books, or it could be as simple as doing a web search to get quick answers. At this stage, you don’t necessarily need to prove or disprove your hypothesis, you merely need to gather the information that you require in order to prove or disprove it yourself. You must know first if you are right in proving or disproving this hypothesis.

  • Define your variables

Once you have decided what your hypothesis will be, you will have to select variables that are independent and which are dependent. Always keep in mind that independent variables are the factors that you have absolute control over. Therefore, consider the limits of your experiment before you finalize your hypothesis. 

  • Phrase it as an if-then statement

While you write your hypothesis, phrase it by using an if-then format, for example, “ If  I water a plant every day,  then  it will grow better.” This format may be tricky while dealing with multiple variables, however, generally, it’s a reliable way of expressing the cause-and-effect relationship that you’re testing. 

  • Collect data to support your hypothesis

A hypothesis is a means to an end. The prime concern of any scientific research is the conclusion. Once you are done with all 4 steps, you can start with your experiments. You will be gathering data to support your hypothesis. However, if your research turns out to prove it wrong, then don’t worry, as all this is a part of the scientific method. 

  • Write with confidence

Lastly, you would want a record of your findings in a research paper for other people to read. Write the research paper and be proud that you made it to the last step.

a person sitting in front of a laptop

To Conclude

  • The hypothesis is a tentative explanation needing scientific testing.
  • Scientific Theory is backed by evidence. It is an extensively tested explanation.
  • The hypothesis is untested. On the other hand, the Theory is based on existing data.
  • Hypotheses need more testing. And theories encompass broad data.
  • Scientific theories change, e.g., the Big Bang and the Theory of Evolution.
  • The hypothesis can be defined as an “educated guess” to be tested.
  • Constructing a hypothesis involves questioning, researching, defining variables, and gathering data.
  • Hypotheses explain your predictions. And the theories explain natural aspects.

Other Articles

  • Fahrenheit and Celsius: Differences Explained
  • Digital vs. Electronic (What’s the Difference?)
  • What Is the Difference Between Sciatica and Meralgia Paresthetica? (Explained)
  • Magnetic & Electric Fields’ Phase Difference

Library homepage

  • school Campus Bookshelves
  • menu_book Bookshelves
  • perm_media Learning Objects
  • login Login
  • how_to_reg Request Instructor Account
  • hub Instructor Commons
  • Download Page (PDF)
  • Download Full Book (PDF)
  • Periodic Table
  • Physics Constants
  • Scientific Calculator
  • Reference & Cite
  • Tools expand_more
  • Readability

selected template will load here

This action is not available.

Chemistry LibreTexts

1.1: Hypothesis, Theories, and Laws

  • Last updated
  • Save as PDF
  • Page ID 338933

  Learning Objectives

  • Describe the difference between hypothesis and theory as scientific terms.
  • Describe the difference between a theory and scientific law.

Although many have taken science classes throughout the course of their studies, people often have incorrect or misleading ideas about some of the most important and basic principles in science. Most students have heard of hypotheses, theories, and laws, but what do these terms really mean? Prior to reading this section, consider what you have learned about these terms before. What do these terms mean to you? What do you read that contradicts or supports what you thought?

What is a Fact?

A fact is a basic statement established by experiment or observation. All facts are true under the specific conditions of the observation.

What is a Hypothesis?

One of the most common terms used in science classes is a "hypothesis". The word can have many different definitions, depending on the context in which it is being used:

  • An educated guess: a scientific hypothesis provides a suggested solution based on evidence.
  • Prediction: if you have ever carried out a science experiment, you probably made this type of hypothesis when you predicted the outcome of your experiment.
  • Tentative or proposed explanation: hypotheses can be suggestions about why something is observed. In order for it to be scientific, however, a scientist must be able to test the explanation to see if it works and if it is able to correctly predict what will happen in a situation. For example, "if my hypothesis is correct, we should see ___ result when we perform ___ test."
A hypothesis is very tentative; it can be easily changed.

What is a Theory?

The United States National Academy of Sciences describes what a theory is as follows:

"Some scientific explanations are so well established that no new evidence is likely to alter them. The explanation becomes a scientific theory. In everyday language a theory means a hunch or speculation. Not so in science. In science, the word theory refers to a comprehensive explanation of an important feature of nature supported by facts gathered over time. Theories also allow scientists to make predictions about as yet unobserved phenomena."

"A scientific theory is a well-substantiated explanation of some aspect of the natural world, based on a body of facts that have been repeatedly confirmed through observation and experimentation. Such fact-supported theories are not "guesses" but reliable accounts of the real world. The theory of biological evolution is more than "just a theory." It is as factual an explanation of the universe as the atomic theory of matter (stating that everything is made of atoms) or the germ theory of disease (which states that many diseases are caused by germs). Our understanding of gravity is still a work in progress. But the phenomenon of gravity, like evolution, is an accepted fact.

Note some key features of theories that are important to understand from this description:

  • Theories are explanations of natural phenomena. They aren't predictions (although we may use theories to make predictions). They are explanations as to why we observe something.
  • Theories aren't likely to change. They have a large amount of support and are able to satisfactorily explain numerous observations. Theories can, indeed, be facts. Theories can change, but it is a long and difficult process. In order for a theory to change, there must be many observations or pieces of evidence that the theory cannot explain.
  • Theories are not guesses. The phrase "just a theory" has no room in science. To be a scientific theory carries a lot of weight; it is not just one person's idea about something
Theories aren't likely to change.

What is a Law?

Scientific laws are similar to scientific theories in that they are principles that can be used to predict the behavior of the natural world. Both scientific laws and scientific theories are typically well-supported by observations and/or experimental evidence. Usually scientific laws refer to rules for how nature will behave under certain conditions, frequently written as an equation. Scientific theories are more overarching explanations of how nature works and why it exhibits certain characteristics. As a comparison, theories explain why we observe what we do and laws describe what happens.

For example, around the year 1800, Jacques Charles and other scientists were working with gases to, among other reasons, improve the design of the hot air balloon. These scientists found, after many, many tests, that certain patterns existed in the observations on gas behavior. If the temperature of the gas is increased, the volume of the gas increased. This is known as a natural law. A law is a relationship that exists between variables in a group of data. Laws describe the patterns we see in large amounts of data, but do not describe why the patterns exist.

What is a Belief?

A belief is a statement that is not scientifically provable. Beliefs may or may not be incorrect; they just are outside the realm of science to explore.

Laws vs. Theories

A common misconception is that scientific theories are rudimentary ideas that will eventually graduate into scientific laws when enough data and evidence has accumulated. A theory does not change into a scientific law with the accumulation of new or better evidence. Remember, theories are explanations and laws are patterns we see in large amounts of data, frequently written as an equation. A theory will always remain a theory; a law will always remain a law.

Video \(\PageIndex{1}\): What’s the difference between a scientific law and theory?

  • A hypothesis is a tentative explanation that can be tested by further investigation.
  • A theory is a well-supported explanation of observations.
  • A scientific law is a statement that summarizes the relationship between variables.
  • An experiment is a controlled method of testing a hypothesis.

Contributions & Attributions

Marisa Alviar-Agnew  ( Sacramento City College )

Henry Agnew (UC Davis)

hypothesis and scientific theory

The Simulation Hypothesis & Whether or Not We’re Living It

Ever wondered if the world around you is nothing more than an elaborate illusion?

The modern twist on this age-old query asks whether our universe resembles a computer game, with us as mere characters following a pre-written script. Popularized by the “Matrix” series, this theory suggests what we perceive as reality could be a sophisticated simulation.

But is such a concept feasible, or is it just the stuff of science fiction?

The Quantum Argument Against Simulated Reality

The idea that life and reality, as we know it, could be computer-generated illusions captivates many. However, a closer look at the laws governing our universe—particularly through the lens of quantum mechanics—presents a compelling counterargument. 1

Photosynthesis, a process central to life on Earth, operates on principles of quantum mechanics that remain not fully understood.

The notion that a flower, through its natural processes, can perform computations beyond the reach of our most advanced quantum computers is a testament to the complexity and sophistication of nature.

This complexity extends to the weather, ecosystems, and even our own bodies, governed by interactions that require an astronomical amount of information to simulate accurately.

Computational Power Limitations

The sheer scale of computational power needed to simulate every atom within a seemingly mundane object, like a goldfish bowl, highlights the limitations of current and foreseeable technology.

The number of atoms and their interactions in just a small portion of our environment is so vast that it defies practical computation. This immense complexity suggests that simulating an entire universe, with all its quantum intricacies, is far beyond our reach.

The interactions of atoms and particles, obeying the principles of quantum mechanics, hold our physical reality together.

Without these principles, the very fabric of our existence would unravel.

Quantum Mechanics in Reality

Quantum mechanics doesn’t just challenge a simulated universe’s feasibility; it underscores our existence’s fundamental nature. The unpredictability and interconnectedness of particles at the quantum level contribute to a reality that is fundamentally irreducible to binary computation.

This unpredictability extends to the concept of parallel universes, further complicating the notion of simulating an entire universe with fidelity.

With its infinite possibilities, the quantum realm highlights the limitations inherent in the simulation hypothesis.

The Butterfly Effect & Computational Complexity

The “butterfly effect” illustrates the sensitivity of complex systems to initial conditions, where minor changes can lead to unforeseeable outcomes. While popular in discussions about chaos theory, this concept poses a significant challenge to simulating a universe.

The computational power required to account for every variable, down to the flap of a butterfly’s wings, and predict its impact on global weather patterns exemplifies the impracticality of creating a fully simulated reality. 2

The Almost Simulation: A Middle Ground?

Acknowledging the limitations doesn’t mean dismissing the broader implications of simulation theories. While a perfect, atom-by-atom simulation may be unattainable, the notion of an “almost simulation” introduces a fascinating grey area.

This concept suggests a simulation that captures the essence of macroscopic features of reality without delving into the quantum minutiae. However, even this more limited approach encounters significant computational and philosophical obstacles, leaving the simulation hypothesis more speculative than ever.

The Simulation Is Unlikely

The debate over whether our universe is a grand simulation continues to intrigue and challenge our understanding of reality. While the allure of such a theory is undeniable, the scientific evidence—rooted in the principles of quantum mechanics and the computational limits we face—suggests a universe far too complex and nuanced to be the product of digital fabrication. 

As we push the boundaries of technology and understanding, the mysteries of the cosmos remain a humbling reminder of the wonders and realities beyond our current grasp.

Reality, it seems, is not an illusion but a quantum tapestry woven with so much complexity, unpredictability, and infinite possibility.

Baba Vanga, a blind mystic from Bulgaria, has blown minds with her spot-on predictions, even years after she passed away.

Imagine being 12 and suddenly losing your sight in an accident, but then, getting something superpowers can’t beat: the ability to see the future. 

Here are six of her visions of the future.

  • Don’t Miss: Blind Mystic Leaves Us With 6 Prophetic Visions & Poignant Message Before Her Death
  • mdpi.com/1099-4300/22/2/247
  • plus.maths.org/content/maths-minute-chaos

The Simulation Hypothesis & Whether or Not We’re Living It

IMAGES

  1. Primary Difference Between Hypothesis and Theory

    hypothesis and scientific theory

  2. The Scientific Method

    hypothesis and scientific theory

  3. Scientific Law Definition and Examples

    hypothesis and scientific theory

  4. Scientific hypothesis

    hypothesis and scientific theory

  5. How to Write a Hypothesis

    hypothesis and scientific theory

  6. PPT

    hypothesis and scientific theory

VIDEO

  1. Basics of Hypothesis, theory and scientific laws

  2. Scientific Theory #Gaia #Hypothesis #subscribe

  3. In the scientific method, a hypothesis is an a observation b measurement c test d propos

  4. Hypothesis theory 🌍🪐…. #viral #youtubeshorts #astronomy #science

  5. What Is A Hypothesis?

  6. Research Hypothesis and its Types with examples /urdu/hindi

COMMENTS

  1. Theory vs. Hypothesis: Basics of the Scientific Method

    A scientific hypothesis is a proposed explanation for an observable phenomenon. In other words, a hypothesis is an educated guess about the relationship between multiple variables. A hypothesis is a fresh, unchallenged idea that a scientist proposes prior to conducting research. The purpose of a hypothesis is to provide a tentative explanation ...

  2. Scientific hypothesis

    hypothesis. science. scientific hypothesis, an idea that proposes a tentative explanation about a phenomenon or a narrow set of phenomena observed in the natural world. The two primary features of a scientific hypothesis are falsifiability and testability, which are reflected in an "If…then" statement summarizing the idea and in the ...

  3. Hypothesis vs. Theory: The Difference Explained

    Toward that end, science employs a particular vocabulary for describing how ideas are proposed, tested, and supported or disproven. And that's where we see the difference between a hypothesis and a theory.. A hypothesis is an assumption, something proposed for the sake of argument so that it can be tested to see if it might be true.. In the scientific method, the hypothesis is constructed ...

  4. Scientific Hypothesis, Theory, Law Definitions

    A scientific theory summarizes a hypothesis or group of hypotheses that have been supported with repeated testing. A theory is valid as long as there is no evidence to dispute it. Therefore, theories can be disproven. Basically, if evidence accumulates to support a hypothesis, then the hypothesis can become accepted as a good explanation of a ...

  5. Scientific Theory Definition and Examples

    A scientific theory is a well-established explanation of some aspect of the natural world. Theories come from scientific data and multiple experiments. While it is not possible to prove a theory, a single contrary result using the scientific method can disprove it. ... Theory vs Hypothesis. A hypothesis is a proposition that is tested via an ...

  6. Hypothesis vs Theory

    A hypothesis is either a suggested explanation for an observable phenomenon, or a reasoned prediction of a possible causal correlation among multiple phenomena. In science, a theory is a tested, well-substantiated, unifying explanation for a set of verified, proven factors. A theory is always backed by evidence; a hypothesis is only a suggested possible outcome, and is testable and falsifiable.

  7. Scientific theory

    scientific theory, systematic ideational structure of broad scope, conceived by the human imagination, that encompasses a family of empirical (experiential) laws regarding regularities existing in objects and events, both observed and posited. A scientific theory is a structure suggested by these laws and is devised to explain them in a scientifically rational manner.

  8. Scientific hypothesis, theory, and model explained

    Scientific model is used to test the scientific hypothesis or to provide a representation of a scientific theory. In the case of plate tectonics, scientists came up with a hypothesis, an idea that Earth's crust was divided into plates that can move or shift. And then models were developed to simulate, or represent, the plates on Earth's crust.

  9. The scientific method (article)

    The scientific method. At the core of biology and other sciences lies a problem-solving approach called the scientific method. The scientific method has five basic steps, plus one feedback step: Make an observation. Ask a question. Form a hypothesis, or testable explanation. Make a prediction based on the hypothesis.

  10. On the scope of scientific hypotheses

    2. The scientific hypothesis. In this section, we will describe a functional and descriptive role regarding how scientists use hypotheses. Jeong & Kwon [] investigated and summarized the different uses the concept of 'hypothesis' had in philosophical and scientific texts.They identified five meanings: assumption, tentative explanation, tentative cause, tentative law, and prediction.

  11. Scientific Hypotheses: Writing, Promoting, and Predicting Implications

    One of the experts in the field defines "hypothesis" as a well-argued analysis of available evidence to provide a realistic (scientific) explanation of existing facts, fill gaps in public understanding of sophisticated processes, and propose a new theory or a test.4 A hypothesis can be proven wrong partially or entirely. However, even such ...

  12. 1.5: Hypothesis, Theories, and Laws

    Henry Agnew (UC Davis) 1.5: Hypothesis, Theories, and Laws is shared under a not declared license and was authored, remixed, and/or curated by LibreTexts. Although all of us have taken science classes throughout the course of our study, many people have incorrect or misleading ideas about some of the most important and basic principles in science.

  13. The Structure of Scientific Theories

    The Structure of Scientific Theories. First published Thu Mar 5, 2015; substantive revision Tue Dec 29, 2020. Scientific inquiry has led to immense explanatory and technological successes, partly as a result of the pervasiveness of scientific theories. Relativity theory, evolutionary theory, and plate tectonics were, and continue to be, wildly ...

  14. 1.3: Hypothesis, Theories, and Laws

    Henry Agnew (UC Davis) 1.3: Hypothesis, Theories, and Laws is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts. Although all of us have taken science classes throughout the course of our study, many people have incorrect or misleading ideas about some of the most important and basic principles in ...

  15. Scientific theory

    A scientific theory is an explanation of an aspect of the natural world and universe that can be (or a fortiori, that has been) repeatedly tested and corroborated in accordance with the scientific method, using accepted protocols of observation, measurement, and evaluation of results. Where possible, some theories are tested under controlled ...

  16. Hypothesis, Model, Theory, and Law

    Scientific Hypothesis, Model, Theory, and Law. By Anne Marie Helmenstine, Ph.D. The word "law" is often invoked in reference to a specific mathematical equation that relates the different elements within a theory. Pascal's Law refers an equation that describes differences in pressure based on height. In the overall theory of universal ...

  17. What is a scientific theory?

    A scientific theory is a structured explanation to explain a group of facts or phenomena in the natural world that often incorporates a scientific hypothesis and scientific laws. The scientific ...

  18. 1.2: Theories, Hypotheses and Models

    A "hypothesis" is a consequence of the theory that one can test. From Chloë's Theory, we have the hypothesis that an object will take 2-√ 2 times longer to fall from 1m 1 m than from 2 m 2 m. We can formulate the hypothesis based on the theory and then test that hypothesis. If the hypothesis is found to be invalidated by experiment ...

  19. Hypothesis vs. Theory: Understanding Scientific Concepts

    Both a hypothesis and a theory are fundamental to scientific progress, but they serve different purposes. A hypothesis proposes a tentative explanation that can be tested and falsified, while a theory summarizes a hypothesis, or a group of hypotheses, that have been supported with repeated testing and is backed by substantial evidence.

  20. The Difference Between a Scientific Hypothesis, Theory, and Law

    A scientific theory can be defined as a series of repeatedly tested and verified hypotheses and concepts. Scientific theories are based on hypotheses that are constructed and tested using the ...

  21. Hypothesis VS A Scientific Theory: A Comparison

    A hypothesis' is basically an educational guess. A hypothesis is often tentative, an assumption or suggestion that is formed strictly for the sole objective of being tested. A hypothesis is an explanation for a phenomenon. A hypothesis is based on prior observations that couldn't be explained with the scientific theories already available.

  22. 1.1: Hypothesis, Theories, and Laws

    Henry Agnew (UC Davis) 1.1: Hypothesis, Theories, and Laws is shared under a not declared license and was authored, remixed, and/or curated by LibreTexts. Although all of us have taken science classes throughout the course of our study, many people have incorrect or misleading ideas about some of the most important and basic principles in science.

  23. Physicist Claims Universe Has No Dark Matter And Is 27 Billion Years

    The consequences of Gupta's version of the tired light hypothesis - what is referred to as covarying coupling constants plus tired light, or CCC+TL - would affect the Universe expansion, doing away with mysterious pushing forces of dark energy and blaming changing interactions between known particles for the increased stretching of space.

  24. The Simulation Hypothesis & Whether or Not We're Living It

    While the allure of such a theory is undeniable, the scientific evidence—rooted in the principles of quantum mechanics and the computational limits we face—suggests a universe far too complex ...