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

essay about philosophy of science

Understanding Science

How science REALLY works...

  • The philosophy of science is a field that deals with what science is, how it works, and the logic through which we build scientific knowledge.
  • In this website, we present a rough synthesis of some new and some old ideas from the philosophy of science.

The philosophy of science

In this website, we use a practical checklist to get a basic picture of what ​​ science  is and a flexible flowchart to depict how science works. For most everyday purposes, this gives us a fairly complete picture of what science is and is not. However, there is an entire field of rigorous academic study that deals specifically with what science is, how it works, and the logic through which we build scientific knowledge. This branch of philosophy is handily called the philosophy of science. Many of the ideas that we present in this website are a rough synthesis of some new and some old ideas from the philosophy of science.

Despite its straightforward name, the field is complex and remains an area of current inquiry. Philosophers of science actively study such questions as:

  • What is a ​​ law  of nature? Are there any in non-physical sciences like biology and psychology?
  • What kind of ​​ data  can be used to distinguish between real causes and accidental regularities?
  • How much ​​ evidence  and what kinds of evidence do we need before we accept ​​ hypotheses ?
  • Why do scientists continue to rely on ​​ models  and ​​ theories  which they know are at least partially inaccurate (like Newton’s physics)?

Though they might seem elementary, these questions are actually quite difficult to answer satisfactorily. Opinions on such issues vary widely within the field (and occasionally part ways with the views of scientists themselves — who mainly spend their time  doing  science, not analyzing it abstractly). Despite this diversity of opinion, philosophers of science can largely agree on one thing: there is no single, simple way to define science!

Though the field is highly specialized, a few touchstone ideas have made their way into the mainstream. Here’s a quick explanation of just a few concepts associated with the philosophy of science, which you might (or might not) have encountered.

  • Epistemology  — branch of philosophy that deals with what knowledge is, how we come to ​​ accept  some things as true, and how we justify that acceptance.
  • Empiricism  — set of philosophical approaches to building knowledge that emphasizes the importance of ​​ observable  evidence from the ​​ natural world .
  • Induction  — method of reasoning in which a generalization is argued to be true based on individual examples that seem to fit with that generalization. For example, after observing that trees, bacteria, sea anemones, fruit flies, and humans have cells, one might  inductively  ​​ infer  that all organisms have cells.
  • Deduction  — method of reasoning in which a conclusion is logically reached from premises. For example, if we know the current relative positions of the moon, sun, and Earth, as well as exactly how these move with respect to one another, we can ​​ deduce  the date and location of the next solar eclipse.
  • Parsimony/Occam’s razor  — idea that, all other things being equal, we should prefer a simpler explanation over a more complex one.
  • Demarcation problem  — the problem of reliably distinguishing science from non-science. Modern philosophers of science largely agree that there is no single, simple criterion that can be used to demarcate the boundaries of science.
  • Falsification  — the view, associated with philosopher Karl Popper, that evidence can only be used to rule out ideas, not to support them. Popper proposed that scientific ideas can only be ​​ tested  through ​​ falsification , never through a search for supporting evidence.
  • Paradigm shifts and scientific revolutions  — a view of science, associated with philosopher Thomas Kuhn, which suggests that the history of science can be divided up into times of normal science (when scientists add to, elaborate on, and work with a central, accepted scientific theory) and briefer periods of revolutionary science. Kuhn asserted that during times of revolutionary science, anomalies refuting the accepted theory have built up to such a point that the old theory is broken down and a new one is built to take its place in a so-called “paradigm shift.”

Who’s who in the philosophy of science

If you’re interested in learning more about the philosophy of science, you might want to begin your investigation with some of the big names in the field:

Aristotle (384-322 BC) — Arguably the founder of both science and philosophy of science. He wrote extensively about the topics we now call physics, astronomy, psychology, biology, and chemistry, as well as logic, mathematics, and epistemology.

Francis Bacon (1561-1626) — Promoted a scientific method in which scientists gather many ​​ facts  from observations and ​​ experiments , and then make ​​ inductive inferences  about patterns in nature.

Rene Descartes (1596-1650) — Mathematician, scientist, and philosopher who promoted a scientific method that emphasized deduction from first principles. These ideas, as well as his mathematics, influenced Newton and other figures of the Scientific Revolution.

Piere Duhem (1861-1916) — Physicist and philosopher who defended an extreme form of empiricism. He argued that we cannot draw conclusions about the existence of unobservable entities conjectured by our theories such as atoms and molecules.

Carl Hempel (1905-1997) — Developed influential theories of scientific explanation and theory confirmation. He argued that a phenomenon is “explained” when we can see that it is the logical consequence of a law of nature. He championed a hypothetico-deductive account of confirmation, similar to the way we characterize “making a ​​ scientific argument ” in this website.

Karl Popper (1924-1994) — Argued that falsifiability is both the hallmark of scientific theories and the proper methodology for scientists to employ. He believed that scientists should always regard their theories with a skeptical eye, seeking every opportunity to try to falsify them.

Thomas Kuhn (1922-1996) — Historian and philosopher who argued that the picture of science developed by logical empiricists such as Popper didn’t resemble the history of science. Kuhn famously distinguished between normal science, where scientists solve puzzles within a particular framework or paradigm, and revolutionary science, when the paradigm gets overturned.

Paul Feyerabend (1924-1994) — A rebel within the philosophy of science. He argued that there is no scientific method or, in his words, “anything goes.” Without regard to rational guidelines, scientists do whatever they need to in order to come up with new ideas and persuade others to accept them.

Evelyn Fox Keller (1936-) — Physicist, historian, and one of the pioneers of feminist philosophy of science, exemplified in her study of Barbara McClintock and the history of genetics in the 20th century.

Elliott Sober (1948-) — Known for his work on ​​ parsimony  and the conceptual foundations of evolutionary biology. He is also an important contributor to the biological theory of group selection.

Nancy Cartwright (1944-) — Philosopher of physics known for her claim that the laws of physics “lie” — i.e., that the laws of physics only apply in highly idealized circumstances. She has also worked on causation, interpretations of probability and quantum mechanics, and the metaphysical foundations of modern science.

  • Take a sidetrip

Learn about specialized topics in the philosophy of science with the  Stanford Encyclopedia of Philosophy .

Source material: Godfrey-Smith, P. 2003. Theory and Reality. Chicago: The University of Chicago Press.

Subscribe to our newsletter

  • Understanding Science 101
  • The science flowchart
  • Science stories
  • Grade-level teaching guides
  • Teaching resource database
  • Journaling tool
  • Misconceptions
  • Undergraduate
  • High School
  • Architecture
  • American History
  • Asian History
  • Antique Literature
  • American Literature
  • Asian Literature
  • Classic English Literature
  • World Literature
  • Creative Writing
  • Linguistics
  • Criminal Justice
  • Legal Issues
  • Anthropology
  • Archaeology
  • Political Science
  • World Affairs
  • African-American Studies
  • East European Studies
  • Latin-American Studies
  • Native-American Studies
  • West European Studies
  • Family and Consumer Science
  • Social Issues
  • Women and Gender Studies
  • Social Work
  • Natural Sciences
  • Pharmacology
  • Earth science
  • Agriculture
  • Agricultural Studies
  • Computer Science
  • IT Management
  • Mathematics
  • Investments
  • Engineering and Technology
  • Engineering
  • Aeronautics
  • Medicine and Health
  • Alternative Medicine
  • Communications and Media
  • Advertising
  • Communication Strategies
  • Public Relations
  • Educational Theories
  • Teacher's Career
  • Chicago/Turabian
  • Company Analysis
  • Education Theories
  • Shakespeare
  • Canadian Studies
  • Food Safety
  • Relation of Global Warming and Extreme Weather Condition
  • Movie Review
  • Admission Essay
  • Annotated Bibliography
  • Application Essay
  • Article Critique
  • Article Review
  • Article Writing
  • Book Review
  • Business Plan
  • Business Proposal
  • Capstone Project
  • Cover Letter
  • Creative Essay
  • Dissertation
  • Dissertation - Abstract
  • Dissertation - Conclusion
  • Dissertation - Discussion
  • Dissertation - Hypothesis
  • Dissertation - Introduction
  • Dissertation - Literature
  • Dissertation - Methodology
  • Dissertation - Results
  • GCSE Coursework
  • Grant Proposal
  • Marketing Plan
  • Multiple Choice Quiz
  • Personal Statement
  • Power Point Presentation
  • Power Point Presentation With Speaker Notes
  • Questionnaire
  • Reaction Paper
  • Research Paper
  • Research Proposal
  • SWOT analysis
  • Thesis Paper
  • Online Quiz
  • Literature Review
  • Movie Analysis
  • Statistics problem
  • Math Problem
  • All papers examples
  • How It Works
  • Money Back Policy
  • Terms of Use
  • Privacy Policy
  • We Are Hiring

Philosophy of Science, Essay Example

Pages: 4

Words: 1022

Hire a Writer for Custom Essay

Use 10% Off Discount: "custom10" in 1 Click 👇

You are free to use it as an inspiration or a source for your own work.

Science is the understanding of the philosophy behind the hypothesis of an experiment and being able to articulate it according to the comprehension of data as applied to a conclusion of the results achieved. Through the embracement of psychology, medicine, artificial intelligence, linguistics and other studies of science researchers began to make different interpretations of how computational procedures and deductive reasoning theories could be used to hypothesize how different results could be achieved through the scientific method. Cognitive science methodology has helped to develop different theories based on research of the mind and intelligence using deductive reasoning and picture reasoning theory practise. Linguistic practise methodology has been used to determine how subtle differences take place in the grammatical human language theories. Neuroscientists assist with these studies in relation to how these studies affect the relation to the brain. Neuroscientists are often looking for mental imagery and brain imagery facets with relation to medicinal applications to science. Often medicinal science and psychology study the relation of situations together to form one theory of application.

Theoretically the principles of science are built upon the realities of what we call the laws of science of how we interpret them to be. A regularity of science is simply a general fact that can be proven. If we hypothesize the sky is red and we can prove it by scientific methodology then it is a known fact. The laws of regularities involve minimalization and instances of science. The law of minimalist is the observation of something occurring regularly over a period of time. These regularities can be fined tuned but they are not legal laws but they are scientific laws. There certainly are laws of science such as the gas laws that are used scientifically to calculate other scientific data within the scientific field such as volume expansion and other areas. There should not be the premise to exclude an idea from scientific law simply because there were only a few instances of its occurrence. The gas law is a functional law that is used to function as filling in the gaps of any extraneous situations that may give rise to outside variables within an experiment.

Math and science have a connection since the great philosophers of Plato and Aristotle who brought math and science together. Many were trained first as mathematicians or engineers. The basis is the urge to build and understand our understanding of the world through the eyes of science and precision of mathematics. Scientists have always tried to find out what was so special about mathematics and how it could be applied to science to improve science and its methods and procedures.

Professor Hillary Putnam of Harvard University feels since the 17 th century there has been a decline in the world view based on religion and an up rise on the view based on science and mathematics. Philosophy is built on how things once fit and how they fit now and into the future. The foundations of physics play a large role with mathematics and physics as Newton, Maxwell and Dalton once studied and spoke of. There is an accumulation of learned knowledge and mathematical and physics knowledge to apply to scientific hypothesis and experimentation as related to cultural and social application through history of science. Inductive method logic and methodology was used explicitly to learn new inductive methods for testing throughout the years until now. Matter and motion was the original thought of what we are mad of for over 200 years. Now we know that there is more to the thoughts and exposition of science. We do know that scientific methods help us to completely understand many facets of the world.

The facts of what it or is not true depends of what the human minds believes something to be-vision! What we think of with a very simple transaction within the world is based on our conscious and unconscious interpretation of the world. Perhaps that is why interpretations of science are quite different today as they were many years ago as observed by scientists. We see the world in different categories and organize our ideas and facts within our own minds. In science, sometimes these interpretations that we call them ‘proven facts’ are falsified in a sense because they can be interpreted in two different ways.

After Newton it was thought he had discovered objective laws of nature that were evident of his discoveries called the ‘truth’. Later people began to realize these interpretations were only theories and could be disproven. Science is an ‘objective interpretation’ from observing the world in the eyes of the beholder and it is through various observers producing the same experiment that we get congruent and valid data.Subjective truth is only opinion and objective truth is proven through testing of hypothesis. Quantum mechanics is an example of objective truth where the philosophy of mathematics and physics can be put to work to test a hypothetical theory. The idea that science is objective and always changing does mean it has some fit in this world. Scientist still do put rockets in the air that work. Most of science is built on objective theory but there still exists some subjective theory. The human person will always have some subjective influence on the philosophy of science and mathematics. Religion will affect some theories of science such as the evolution theories. The scientists believe everything is fallible and we will never get to the end of learning all there is to but we can definitively test certain theories along the way to prove certain existence and theories thereof.

So with that said philosophy is a series of testing through objective means to quantify mathematical applications necessary to test a set of hypotheses and theories. Science is not a law in itself and even if it was some laws are discretionary or subjective. Through experimentation and testing by the laws of science we can hope to gain useful knowledge of our past existence, learn innovative ways to produce electricity, supply our own gasoline and oil, make better food and water supply, find cures for diseases and make this world a safer and happier place to live in.

Stuck with your Essay?

Get in touch with one of our experts for instant help!

Historical Lives of Jesus and Mohammed, Essay Example

Debate, Essay Example

Time is precious

don’t waste it!

Plagiarism-free guarantee

Privacy guarantee

Secure checkout

Money back guarantee

E-book

Related Essay Samples & Examples

Voting as a civic responsibility, essay example.

Pages: 1

Words: 287

Utilitarianism and Its Applications, Essay Example

Words: 356

The Age-Related Changes of the Older Person, Essay Example

Pages: 2

Words: 448

The Problems ESOL Teachers Face, Essay Example

Pages: 8

Words: 2293

Should English Be the Primary Language? Essay Example

Words: 999

The Term “Social Construction of Reality”, Essay Example

Words: 371

essay about philosophy of science

Science must become attuned to the subtle conversations that pervade all life, from the primordial to the present

David Waltner-Toews

essay about philosophy of science

History of science

The missing conversation

To the detriment of the public, scientists and historians don’t engage with one another. They must begin a new dialogue

Lorraine Daston & Peter Harrison

essay about philosophy of science

When aggression is viewed as brilliance, it hurts women in science, and science itself

essay about philosophy of science

Philosophy of science

Why not scientism?

Science is not the only form of knowledge but it is the best, being the most successful epistemic enterprise in history

Moti Mizrahi

essay about philosophy of science

Machina mundi

How medieval thinkers foreshadowed modern physics in investigating the character of machines, devices and forces

Henrik Lagerlund & Sylvain Roudaut

essay about philosophy of science

Thinkers and theories

Bernard Williams on Descartes’s audacious endeavour to prove knowledge is possible

essay about philosophy of science

A sliver of reality

Science and mathematics may never fully capture the physical universe. Are there hard limits to human intelligence?

David H Wolpert

essay about philosophy of science

The lady vanishes

The history of ideas still struggles to remember the names of notable women philosophers. Mary Hesse is a salient example

Ann-Sophie Barwich

essay about philosophy of science

What is a law of nature?

Laws of nature are impossible to break, and nearly as difficult to define. Just what kind of necessity do they possess?

essay about philosophy of science

Logic and probability

The unresolved probability paradox that goes to the heart of scientific objectivity

essay about philosophy of science

Quantum theory

Mind-bending new quantum experiments are blurring past, present and future

essay about philosophy of science

Fringe theories stack

Believe in the Loch Ness monster and you’re more likely to believe the Apollo missions were fake. How do weird beliefs work?

Michael D Gordin

essay about philosophy of science

Calculate but don’t shut up

The cliché has it that the Copenhagen interpretation demands adherence without deep enquiry. That does physics a disservice

Jim Baggott

essay about philosophy of science

The beautiful experiment

Science has become extraordinarily technocratic and complex. Is the simple and decisive experiment still a worthy ideal?

Milena Ivanova

essay about philosophy of science

Future of technology

Learn from machine learning

The world is a black box full of extreme specificity: it might be predictable but that doesn’t mean it is understandable

David Weinberger

essay about philosophy of science

Why simplicity works

Does the existence of a multiverse hold the key for why nature’s laws seem so simple?

Johnjoe McFadden

essay about philosophy of science

An idea with bite

The ‘selfish gene’ persists for the reason all good scientific metaphors do: it remains a sharp tool for clear thinking

J Arvid Ågren

essay about philosophy of science

Building ‘bigger and better’ has pushed cosmology forward. Can it take it any further?

essay about philosophy of science

A non-Standard model

Most cosmologists say dark matter must exist. So far, it’s nowhere to be found. A widely scorned rival theory explains why

David Merritt

essay about philosophy of science

Earth science and climate

How much can science really tell us about the future of climate change?

essay about philosophy of science

Mathematics

Against ‘beauty’ in science – how striving for elegance stifles progress

essay about philosophy of science

‘Moving paintings’ evoke a quantum particle collision at the Large Hadron Collider

essay about philosophy of science

Keep science irrational

Is hard data the only path to scientific truth? That’s an absurd, illogical and profoundly useful fiction

Michael Strevens

essay about philosophy of science

Life with purpose

Biologists balk at any talk of ‘goals’ or ‘intentions’ – but a bold new research agenda has put agency back on the table

Philip Ball

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

Einstein’s Philosophy of Science

Albert Einstein (1879–1955) is well known as the most prominent physicist of the twentieth century. His contributions to twentieth-century philosophy of science, though of comparable importance, are less well known. Einstein’s own philosophy of science is an original synthesis of elements drawn from sources as diverse as neo-Kantianism, conventionalism, and logical empiricism, its distinctive feature being its novel blending of realism with a holist, underdeterminationist form of conventionalism. Of special note is the manner in which Einstein’s philosophical thinking was driven by and contributed to the solution of problems first encountered in his work in physics. Equally significant are Einstein’s relations with and influence on other prominent twentieth-century philosophers of science, including Moritz Schlick, Hans Reichenbach, Ernst Cassirer, Philipp Frank, Henri Bergson, Émile Meyerson.

1. Introduction: Was Einstein an Epistemological “Opportunist”?

2. theoretical holism: the nature and role of conventions in science, 3. simplicity and theory choice, 4. univocalness in the theoretical representation of nature, 5. realism and separability, 6. the principle theories—constructive theories distinction, 7. conclusion: albert einstein: philosopher-physicist, einstein’s work, related literature, other internet resources, related entries.

Late in 1944, Albert Einstein received a letter from Robert Thornton, a young African-American philosopher of science who had just finished his Ph.D. under Herbert Feigl at Minnesota and was beginning a new job teaching physics at the University of Puerto Rico, Mayaguez. He had written to solicit from Einstein a few supportive words on behalf of his efforts to introduce “as much of the philosophy of science as possible” into the modern physics course that he was to teach the following spring (Thornton to Einstein, 28 November 1944, EA 61–573). Here is what Einstein offered in reply:

I fully agree with you about the significance and educational value of methodology as well as history and philosophy of science. So many people today—and even professional scientists—seem to me like somebody who has seen thousands of trees but has never seen a forest. A knowledge of the historic and philosophical background gives that kind of independence from prejudices of his generation from which most scientists are suffering. This independence created by philosophical insight is—in my opinion—the mark of distinction between a mere artisan or specialist and a real seeker after truth. (Einstein to Thornton, 7 December 1944, EA 61–574)

That Einstein meant what he said about the relevance of philosophy to physics is evidenced by the fact that he had been saying more or less the same thing for decades. Thus, in a 1916 memorial note for Ernst Mach, a physicist and philosopher to whom Einstein owed a special debt, he wrote:

How does it happen that a properly endowed natural scientist comes to concern himself with epistemology? Is there no more valuable work in his specialty? I hear many of my colleagues saying, and I sense it from many more, that they feel this way. I cannot share this sentiment. When I think about the ablest students whom I have encountered in my teaching, that is, those who distinguish themselves by their independence of judgment and not merely their quick-wittedness, I can affirm that they had a vigorous interest in epistemology. They happily began discussions about the goals and methods of science, and they showed unequivocally, through their tenacity in defending their views, that the subject seemed important to them. Indeed, one should not be surprised at this. (Einstein 1916, 101)

How, exactly, does the philosophical habit of mind provide the physicist with such “independence of judgment”? Einstein goes on to explain:

Concepts that have proven useful in ordering things easily achieve such an authority over us that we forget their earthly origins and accept them as unalterable givens. Thus they come to be stamped as “necessities of thought,” “a priori givens,” etc. The path of scientific advance is often made impassable for a long time through such errors. For that reason, it is by no means an idle game if we become practiced in analyzing the long commonplace concepts and exhibiting those circumstances upon which their justification and usefulness depend, how they have grown up, individually, out of the givens of experience. By this means, their all-too-great authority will be broken. They will be removed if they cannot be properly legitimated, corrected if their correlation with given things be far too superfluous, replaced by others if a new system can be established that we prefer for whatever reason. (Einstein 1916, 102)

One is not surprised at Einstein’s then citing Mach’s critical analysis of the Newtonian conception of absolute space as a paradigm of what Mach, himself, termed the “historical-critical” method of philosophical analysis (Einstein 1916, 101, citing Ch. 2, §§ 6–7 of Mach’s Mechanik , most likely the third edition, Mach 1897).

The place of philosophy in physics was a theme to which Einstein returned time and again, it being clearly an issue of deep importance to him. Sometimes he adopts a modest pose, as in this oft-quoted remark from his 1933 Spencer Lecture:

If you wish to learn from the theoretical physicist anything about the methods which he uses, I would give you the following piece of advice: Don’t listen to his words, examine his achievements. For to the discoverer in that field, the constructions of his imagination appear so necessary and so natural that he is apt to treat them not as the creations of his thoughts but as given realities. (Einstein 1933, 5–6)

More typical, however, is the confident pose he struck three years later in “Physics and Reality”:

It has often been said, and certainly not without justification, that the man of science is a poor philosopher. Why then should it not be the right thing for the physicist to let the philosopher do the philosophizing? Such might indeed be the right thing at a time when the physicist believes he has at his disposal a rigid system of fundamental concepts and fundamental laws which are so well established that waves of doubt can not reach them; but it can not be right at a time when the very foundations of physics itself have become problematic as they are now. At a time like the present, when experience forces us to seek a newer and more solid foundation, the physicist cannot simply surrender to the philosopher the critical contemplation of the theoretical foundations; for, he himself knows best, and feels more surely where the shoe pinches. In looking for a new foundation, he must try to make clear in his own mind just how far the concepts which he uses are justified, and are necessities. (Einstein 1936, 349)

What kind of philosophy might we expect from the philosopher-physicist? One thing that we should not expect from a physicist who takes the philosophical turn in order to help solve fundamental physical problems is a systematic philosophy:

The reciprocal relationship of epistemology and science is of noteworthy kind. They are dependent upon each other. Epistemology without contact with science becomes an empty scheme. Science without epistemology is—insofar as it is thinkable at all—primitive and muddled. However, no sooner has the epistemologist, who is seeking a clear system, fought his way through to such a system, than he is inclined to interpret the thought-content of science in the sense of his system and to reject whatever does not fit into his system. The scientist, however, cannot afford to carry his striving for epistemological systematic that far. He accepts gratefully the epistemological conceptual analysis; but the external conditions, which are set for him by the facts of experience, do not permit him to let himself be too much restricted in the construction of his conceptual world by the adherence to an epistemological system. He therefore must appear to the systematic epistemologist as a type of unscrupulous opportunist: he appears as realist insofar as he seeks to describe a world independent of the acts of perception; as idealist insofar as he looks upon the concepts and theories as free inventions of the human spirit (not logically derivable from what is empirically given); as positivist insofar as he considers his concepts and theories justified only to the extent to which they furnish a logical representation of relations among sensory experiences. He may even appear as Platonist or Pythagorean insofar as he considers the viewpoint of logical simplicity as an indispensable and effective tool of his research. (Einstein 1949, 683–684)

But what strikes the “systematic epistemologist” as mere opportunism might appear otherwise when viewed from the perspective of a physicist engaged, as Einstein himself put it, in “the critical contemplation of the theoretical foundations.” The overarching goal of that critical contemplation was, for Einstein, the creation of a unified foundation for physics after the model of a field theory like general relativity (see Sauer 2014 for non-technical overview on Einstein’s approach to the unified field theory program). Einstein failed in his quest, but there was a consistency and constancy in the striving that informed as well the philosophy of science developing hand in hand with the scientific project.

Indeed, from early to late a few key ideas played the central, leading role in Einstein’s philosophy of science, ideas about which Einstein evinced surprisingly little doubt even while achieving an ever deeper understanding of their implications. For the purposes of the following comparatively brief overview, we can confine our attention to just five topics:

  • Theoretical holism.
  • Simplicity and theory choice.
  • Univocalness in the theoretical representation of nature.
  • Realism and separability.
  • The principle theories-constructive theories distinction.

The emphasis on the continuity and coherence in the development of Einstein’s philosophy of science contrasts with an account such as Gerald Holton’s (1968), which claims to find a major philosophical break in the mid-1910s, in the form of a turn away from a sympathy for an anti-metaphysical positivism and toward a robust scientific realism. Holton sees this turn being driven by Einstein’s alleged realization that general relativity, by contrast with special relativity, requires a realistic ontology. However, Einstein was probably never an ardent “Machian” positivist, [ 1 ] and he was never a scientific realist, at least not in the sense acquired by the term “scientific realist” in later twentieth century philosophical discourse (see Howard 1993). Einstein expected scientific theories to have the proper empirical credentials, but he was no positivist; and he expected scientific theories to give an account of physical reality, but he was no scientific realist. Moreover, in both respects his views remained more or less the same from the beginning to the end of his career.

Why Einstein did not think himself a realist (he said so explicitly) is discussed below. Why he is not to be understood as a positivist deserves a word or two of further discussion here, if only because the belief that he was sympathetic to positivism, at least early in his life, is so widespread (for a fuller discussion, see Howard 1993).

That Einstein later repudiated positivism is beyond doubt. Many remarks from at least the early 1920s through the end of his life make this clear. In 1946 he explained what he took to be Mach’s basic error:

He did not place in the correct light the essentially constructive and speculative nature of all thinking and more especially of scientific thinking; in consequence, he condemned theory precisely at those points where its constructive-speculative character comes to light unmistakably, such as in the kinetic theory of atoms. (Einstein 1946, 21)

Is Einstein here also criticizing his own youthful philosophical indiscretions? The very example that Einstein gives here makes any such interpretation highly implausible, because one of Einstein’s main goals in his early work on Brownian motion (Einstein 1905b) was precisely to prove the reality of atoms, this in the face of the then famous skepticism of thinkers like Mach and Wilhelm Ostwald:

My principal aim in this was to find facts that would guarantee as much as possible the existence of atoms of definite size.… The agreement of these considerations with experience together with Planck’s determination of the true molecular size from the law of radiation (for high temperatures) convinced the skeptics, who were quite numerous at that time (Ostwald, Mach), of the reality of atoms. (Einstein 1946, 45, 47)

Why, then, is the belief in Einstein’s early sympathy for positivism so well entrenched?

The one piece of evidence standardly cited for a youthful flirtation with positivism is Einstein’s critique of the notion of absolute distant simultaneity in his 1905 paper on special relativity (Einstein 1905c). Einstein speaks there of “observers,” but in an epistemologically neutral way that can be replaced by talk of an inertial frame of reference. What really bothers Einstein about distant simultaneity is not that it is observationally inaccessible but that it involves a two-fold arbitrariness, one in the choice of an inertial frame of reference and one in the stipulation within a given frame of a convention regarding the ratio of the times required for a light signal to go from one stationary observer to another and back again. Likewise, Einstein faults classical Maxwellian electrodynamics for an asymmetry in the way it explains electromagnetic induction depending on whether it is the coil or the magnet that is assumed to be at rest. If the effect is the same—a current in the coil—why, asks Einstein, should there be two different explanations: an electrical field created in the vicinity of a moving magnet or an electromotive force induced in a conductor moving through a stationary magnetic field? To be sure, whether it is the coil or the magnet that is taken to be at rest makes no observable difference, but the problem, from Einstein’s point of view, is the asymmetry in the two explanations. Even the young Einstein was no positivist.

First generation logical empiricists sought to legitimate their movement in part by claiming Einstein as a friend. They may be forgiven their putting a forced interpretation on arguments taken out of context. We can do better.

Einstein’s philosophy of science is an original synthesis drawing upon many philosophical resources, from neo-Kantianism to Machian empiricism and Duhemian conventionalism. Other thinkers and movements, most notably the logical empiricists, drew upon the same resources. But Einstein put the pieces together in a manner importantly different from Moritz Schlick, Hans Reichenbach, and Rudolf Carnap, and he argued with them for decades about who was right (however much they obscured these differences in representing Einstein publicly as a friend of logical empiricism and scientific philosophy). Starting from the mid-1920s till the end of the decade Einstein show some interest in the rationalistic realism of Émile Meyerson (Einstein, 1928; cf. Giovanelli 2018; on the contemporary debate between Einstein and Bergson, see Canales 2015). Understanding how Einstein puts those pieces together therefore sheds light not only on the philosophical aspect of his own achievements in physics but also upon the larger history of the development of the philosophy of science in the twentieth century.

Any philosophy of science must include an account of the relation between theory and evidence. Einstein learned about the historicity of scientific concepts from Mach. But his preferred way of modeling the logical relationship between theory and evidence was inspired mainly by his reading of Pierre Duhem’s La Théorie physique: son objet et sa structure (Duhem 1906). Einstein probably first read Duhem, or at least learned the essentials of Duhem’s philosophy of science around the fall of 1909, when, upon returning to Zurich from the patent office in Bern to take up his first academic appointment at the University of Zurich, he became the upstairs neighbor of his old friend and fellow Zurich physics student, Friedrich Adler. Just a few months before, Adler had published the German translation of La Théorie physique (Duhem 1908), and the philosophy of science became a frequent topic of conversation between the new neighbors, Adler and Einstein (see Howard 1990a).

Theoretical holism and the underdetermination of theory choice by empirical evidence are the central theses in Duhem’s philosophy of science. His argument, in brief, is that at least in sciences like physics, where experiment is dense with sophisticated instrumentation whose employment itself requires theoretical interpretation, hypotheses are not tested in isolation but only as part of whole bodies of theory. It follows that when there is a conflict between theory and evidence, the fit can be restored in a multiplicity of different ways. No statement is immune to revision because of a presumed status as a definition or thanks to some other a priori warrant, and most any statement can be retained on pain of suitable adjustments elsewhere in the total body of theory. Hence, theory choice is underdetermined by evidence.

That Einstein’s exposure to Duhem’s philosophy of science soon left its mark is evident from lecture notes that Einstein prepared for a course on electricity and magnetism at the University of Zurich in the winter semester of 1910/11. Einstein asks how one can assign a definite electrical charge everywhere within a material body, if the interior of the body is not accessible to test particles. A “Machian” positivist would deem such direct empirical access necessary for meaningful talk of a charge distribution in the interior of a sold. Einstein argues otherwise:

We have seen how experience led to the introd. of the concept of the quantity of electricity. it was defined by means of the forces that small electrified bodies exert on each other. But now we extend the application of the concept to cases in which this definition cannot be applied directly as soon as we conceive the el. forces as forces exerted on electricity rather than on material particles. We set up a conceptual system the individual parts of which do not correspond directly to empirical facts. Only a certain totality of theoretical material corresponds again to a certain totality of experimental facts. We find that such an el. continuum is always applicable only for the representation of el. states of affairs in the interior of ponderable bodies. Here too we define the vector of el. field strength as the vector of the mech. force exerted on the unit of pos. electr. quantity inside a body. But the force so defined is no longer directly accessible to exp. It is one part of a theoretical construction that can be correct or false, i.e., consistent or not consistent with experience, only as a whole . ( Collected Papers of Albert Einstein , hereafter CPAE, Vol. 3, Doc. 11 [pp. 12–13])

One can hardly ask for a better summary of Duhem’s point of view in application to a specific physical theory. Explicit citations of Duhem by Einstein are rare (for details, see Howard 1990a). But explicit invocations of a holist picture of the structure and empirical interpretation of theories started to prevail at the turn of the 1920s.

During the decade 1905–1915, Einstein had more or less explicitly assumed that in a good theory there are certain individual parts that can be directly coordinated with the behavior of physically-existent objects used as probes. A theory can be said to be ‘true or false’ if such objects respectively behave or do not behave as predicted. In special relativity, as in classical mechanics, the fundamental geometrical/kinematical variables, the space and time coordinates, are measured with rods and clocks separately from the other non-geometrical variables, say, charge electric field strengths, which were supposed to be defined by measuring the force on a charge test particle. In general relativity, coordinates are no longer directly measurable independently from the gravitational field. Still, the line element \(ds\) (distance between nearby spacetime points) was supposed to have a ‘natural’ distance that can be measured with rods and clocks. In the late 1910s, pressed by the epistemological objections raised by different interlocutors—in particular Hermann Weyl (Ryckman 2005) and the young Wolfgang Pauli (Stachel, 2005)—Einstein was forced to recognize that this epistemological model was at most a provisional compromise. In principle rod- and clock-like structures should emerge as solutions of a future relativistic theory of matter, possibly a field theory encompassing gravitation and electromagnetism. In this context, the sharp distinction between rods and clocks that serve to define the geometrical/kinematical structure of the theory and other material systems would become questionable. Einstein regarded such distinction as provisionally necessary, give the current state of physics. However, he recognized that in principle a physical theory should construct rods and clocks as solutions to its equations (see Ryckman 2017, ch. VII for an overview on Einstein view on the relation between geometry and experience).

Einstein addressed this issue in several popular writings during the 1920s, in particular, the famous lecture Geometrie und Erfahrung (Einstein 1921, see also Einstein, 1923, Einstein, 1924, Einstein 1926; Einstein 1926; see Giovanelli 2014 for an overview). Sub specie temporis , he argued, it was useful to compare the geometrical/kinematical structures of the theory with experience separately from the rest of physics. Sub specie aeterni , however, only geometry and physics taken together can be said to be ‘true or false.’ This epistemological model became more appropriate, while Einstein was moving beyond general relativity in the direction of theory unifying the gravitational and the electromagnetic field. Einstein had to rely on progressively more abstract geometrical structures which could not be defined in terms of the behavior of some physical probes. Thus, the use of such structures was justified because of their role in the theory as a whole. In the second half of the 1920s, in correspondence with Reichenbach (Giovanelli 2017) and Meyerson (Giovanelli 2018), Einstein even denied that the very distinction between geometrical and non-geometrical is meaningful (Lehmkuhl 2014).

A different, but especially interesting example of Einstein’s reliance on a form of theoretical holism is found in a review that Einstein wrote in 1924 of Alfred Elsbach’s Kant und Einstein (1924), one of the flood of books and articles then trying to reconcile the Kant’s philosophy. Having asserted that relativity theory is incompatible with Kant’s doctrine of the a priori, Einstein explains why, more generally, he is not sympathetic with Kant:

This does not, at first, preclude one’s holding at least to the Kantian problematic , as, e.g., Cassirer has done. I am even of the opinion that this standpoint can be rigorously refuted by no development of natural science. For one will always be able to say that critical philosophers have until now erred in the establishment of the a priori elements, and one will always be able to establish a system of a priori elements that does not contradict a given physical system. Let me briefly indicate why I do not find this standpoint natural. A physical theory consists of the parts (elements) A, B, C, D, that together constitute a logical whole which correctly connects the pertinent experiments (sense experiences). Then it tends to be the case that the aggregate of fewer than all four elements, e.g., A, B, D, without C, no longer says anything about these experiences, and just as well A, B, C without D. One is then free to regard the aggregate of three of these elements, e.g., A, B, C as a priori, and only D as empirically conditioned. But what remains unsatisfactory in this is always the arbitrariness in the choice of those elements that one designates as a priori, entirely apart from the fact that the theory could one day be replaced by another that replaces certain of these elements (or all four) by others. (Einstein 1924, 1688–1689)

Einstein’s point seems to be that while one can always choose to designate selected elements as a priori and, hence, non-empirical, no principle determines which elements can be so designated, and our ability thus to designate them derives from the fact that it is only the totality of the elements that possesses empirical content.

Much the same point could be made, and was made by Duhem himself (see Duhem 1906, part 2, ch. 6, sects. 8 and 9), against those who would insulate certain statements against empirical refutation by claiming for them the status of conventional definitions. Edouard Le Roy (1901) had argued thus about the law of free fall. It could not be refuted by experiment because it functioned as a definition of “free fall.” And Henri Poincaré (1901) said much the same about the principles of mechanics more generally. As Einstein answered the neo-Kantians, so Duhem answered this species of conventionalist: Yes, experiment cannot refute, say, the law of free fall by itself, but only because it is part of a larger theoretical whole that has empirical content only as a whole, and various other elements of that whole could as well be said to be, alone, immune to refutation.

That Einstein should deploy against the neo-Kantians in the early 1920s the argument that Duhem used against the conventionalism of Poincaré and Le Roy is interesting from the point of view of Einstein’s relationships with those who were leading the development of logical empiricism and scientific philosophy in the 1920s, especially Schlick and Reichenbach. Einstein shared with Schlick and Reichenbach the goal of crafting a new form of empiricism that would be adequate to the task of defending general relativity against neo-Kantian critiques (see Schlick 1917 and 1921, and Reichenbach 1920, 1924, and 1928; for more detail, see Howard 1994a). But while they all agreed that what Kant regarded as the a priori element in scientific cognition was better understood as a conventional moment in science, they were growing to disagree dramatically over the nature and place of conventions in science. The classic logical empiricist view that the moment of convention was restricted to conventional coordinating definitions that endow individual primitive terms, worked well, but did not comport well with the holism about theories

It was this argument over the nature and place of conventions in science that underlies Einstein’s gradual philosophical estrangement from Schlick and Reichenbach in the 1920s. Serious in its own right, the argument over conventions was entangled with two other issues as well, namely, realism and Einstein’s famous view of theories as the “free creations of the human spirit” (see, for example, Einstein 1921). In both instances what troubled Einstein was that a verificationist semantics made the link between theory and experience too strong, leaving too small a role for theory, itself, and the creative theorizing that produces it.

If theory choice is empirically determinate, especially if theoretical concepts are explicitly constructed from empirical primitives, as in Carnap’s program in the Aufbau (Carnap 1928), then it is hard to see how theory gives us a story about anything other than experience. As noted, Einstein was not what we would today call a scientific realist, but he still believed that there was content in theory beyond mere empirical content (on the relations between Einstein’s realism and constructism see Ryckman 2017, ch. 8 and 9). He believed that theoretical science gave us a window on nature itself, even if, in principle, there will be no one uniquely correct story at the level of deep ontology (see below, section 5). And if the only choice in theory choice is one among conventional coordinating definitions, then that is no choice at all, a point stressed by Reichenbach, especially, as an important positive implication of his position. Reichenbach argued that if empirical content is the only content, then empirically equivalent theories have the same content, the difference resulting from their different choices of coordinating definitions being like in kind to the difference between “es regnet” and “il pleut,” or the difference between expressing the result of a measurement in English or metric units, just two different ways of saying the same thing. But then, Einstein would ask, where is there any role for the creative intelligence of the theoretical physicist if there is no room for genuine choice in science, if experience somehow dictates theory construction?

The argument over the nature and role of conventions in science continued to the very end of Einstein’s life, reaching its highest level of sophistication in the exchange between Reichenbach and Einstein the Library of Living Philosopher’s volume, Albert Einstein: Philosopher-Physicist (Schilpp 1949). The question is, again, whether the choice of a geometry is empirical, conventional, or a priori. In his contribution, Reichenbach reasserted his old view that once an appropriate coordinating definition is established, equating some “practically rigid rod” with the geometer’s “rigid body,” then the geometry of physical space is wholly determined by empirical evidence:

The choice of a geometry is arbitrary only so long as no definition of congruence is specified. Once this definition is set up, it becomes an empirical question which geometry holds for physical space.… The conventionalist overlooks the fact that only the incomplete statement of a geometry, in which a reference to the definition of congruence is omitted, is arbitrary. (Reichenbach 1949, 297)

Einstein’s clever reply includes a dialogue between two characters, “Reichenbach” and “Poincaré,” in which “Reichenbach” concedes to “Poincaré” that there are no perfectly rigid bodies in nature and that physics must be used to correct for such things as thermal deformations, from which it follows that what we actually test is geometry plus physics, not geometry alone. Here an “anonymous non-positivist” takes “Poincaré’s” place, out of respect, says Einstein, “for Poincaré’s superiority as thinker and author” (Einstein 1949, 677), but also, perhaps, because he realized that the point of view that follows was more Duhem than Poincaré. The “non-positivist” then argues that one’s granting that geometry and physics are tested together contravenes the positivist identification of meaning with verifiability:

Non-Positivist: If, under the stated circumstances, you hold distance to be a legitimate concept, how then is it with your basic principle (meaning = verifiability)? Must you not come to the point where you deny the meaning of geometrical statements and concede meaning only to the completely developed theory of relativity (which still does not exist at all as a finished product)? Must you not grant that no “meaning” whatsoever, in your sense, belongs to the individual concepts and statements of a physical theory, such meaning belonging instead to the whole system insofar as it makes “intelligible” what is given in experience? Why do the individual concepts that occur in a theory require any separate justification after all, if they are indispensable only within the framework of the logical structure of the theory, and if it is the theory as a whole that stands the test? (Einstein 1949, 678).

Two years before the Quine’s publication of “Two Dogmas of Empiricism” (1951), Einstein here makes explicit the semantic implications of a thoroughgoing holism.

If theory choice is empirically underdetermined, then an obvious question is why we are so little aware of the underdetermination in the day-to-day conduct of science. In a 1918 address celebrating Max Planck’s sixtieth birthday, Einstein approached this question via a distinction between practice and principle:

The supreme task of the physicist is … the search for those most general, elementary laws from which the world picture is to be obtained through pure deduction. No logical path leads to these elementary laws; it is instead just the intuition that rests on an empathic understanding of experience. In this state of methodological uncertainty one can think that arbitrarily many, in themselves equally justified systems of theoretical principles were possible; and this opinion is, in principle , certainly correct. But the development of physics has shown that of all the conceivable theoretical constructions a single one has, at any given time, proved itself unconditionally superior to all others. No one who has really gone deeply into the subject will deny that, in practice, the world of perceptions determines the theoretical system unambiguously, even though no logical path leads from the perceptions to the basic principles of the theory. (Einstein 1918, 31; Howard’s translation)

But why is theory choice, in practice, seemingly empirically determined? Einstein hinted at an answer the year before in a letter to Schlick, where he commended Schlick’s argument that the deep elements of a theoretical ontology have as much claim to the status of the real as do Mach’s elements of sensation (Schlick 1917), but suggested that we are nonetheless speaking of two different kinds of reality. How do they differ?

It appears to me that the word “real” is taken in different senses, according to whether impressions or events, that is to say, states of affairs in the physical sense, are spoken of. If two different peoples pursue physics independently of one another, they will create systems that certainly agree as regards the impressions (“elements” in Mach’s sense). The mental constructions that the two devise for connecting these “elements” can be vastly different. And the two constructions need not agree as regards the “events”; for these surely belong to the conceptual constructions. Certainly on the “elements,” but not the “events,” are real in the sense of being “given unavoidably in experience.” But if we designate as “real” that which we arrange in the space-time-schema, as you have done in the theory of knowledge, then without doubt the “events,” above all, are real.… I would like to recommend a clean conceptual distinction here . (Einstein to Schlick, 21 May 1917, CPAE, Vol. 8, Doc. 343)

Why, in practice, are physicists unaware of underdetermination? It is because ours is not the situation of “two different peoples pursu[ing] physics independently of one another.” Though Einstein does not say it explicitly, the implication seems to be that apparent determination in theory choice is mainly a consequence of our all being similarly socialized as we become members of a common scientific community. Part of what it means to be a member of a such a community is that we have been taught to make our theoretical choices in accord with criteria or values that we hold in common.

For Einstein, as for many others, simplicity is the criterion that mainly steers theory choice in domains where experiment and observation no longer provide an unambiguous guide. This, too, is a theme sounded early and late in Einstein’s philosophical reflections (for more detail, see Howard 1998, Norton 2000, van Dongen 2002, 2010, Giovanelli 2018). For example, the just-quoted remark from 1918 about the apparent determination of theory choice in practice, contrasted with in-principle underdetermination continues:

Furthermore this conceptual system that is univocally coordinated with the world of experience is reducible to a few basic laws from which the whole system can be developed logically. With every new important advance the researcher here sees his expectations surpassed, in that those basic laws are more and more simplified under the press of experience. With astonishment he sees apparent chaos resolved into a sublime order that is to be attributed not to the rule of the individual mind, but to the constitution of the world of experience; this is what Leibniz so happily characterized as “pre-established harmony.” Physicists strenuously reproach many epistemologists for their insufficient appreciation of this circumstance. Herein, it seems to me, lie the roots of the controversy carried on some years ago between Mach and Planck. (Einstein 1918, p. 31)

There is more than a little autobiography here, for as Einstein stressed repeatedly in later years, he understood the success of his own quest for a general theory of relativity as a result of his seeking the simplest set of field equations satisfying a given set of constraints.

Einstein’s celebration of simplicity as a guide to theory choice comes clearly to the fore in the early 1930s, when he was immersed his project of a unified field theory (see, van Dongen 2010 for a reconstruction of the philosophical underpinning of Einstein’s search of a unified field theory). Witness what he wrote in his 1933 Herbert Spencer lecture:

If, then, it is true that the axiomatic foundation of theoretical physics cannot be extracted from experience but must be freely invented, may we ever hope to find the right way? Furthermore, does this right way exist anywhere other than in our illusions? May we hope to be guided safely by experience at all, if there exist theories (such as classical mechanics) which to a large extent do justice to experience, without comprehending the matter in a deep way? To these questions, I answer with complete confidence, that, in my opinion, the right way exists, and that we are capable of finding it. Our experience hitherto justifies us in trusting that nature is the realization of the simplest that is mathematically conceivable. I am convinced that purely mathematical construction enables us to find those concepts and those lawlike connections between them that provide the key to the understanding of natural phenomena. Useful mathematical concepts may well be suggested by experience, but in no way can they be derived from it. Experience naturally remains the sole criterion of the usefulness of a mathematical construction for physics. But the actual creative principle lies in mathematics. Thus, in a certain sense, I take it to be true that pure thought can grasp the real, as the ancients had dreamed. (Einstein 1933, p. 183; Howard’s translation)

Einstein’s conviction that the theoretical physicist must trust simplicity is that his work was moving steadily into domains ever further removed from direct contact with observation and experiment. Einstein started to routinely claim that this was the lesson he had drawn from the way in which he had found general relativity (Norton 2000). There are, however, good reasons to think that Einstein’s selective recollections (Jannsen and Renn 2007) were instrumental to his defense of relying on a purely mathematical strategy in the search for a unified field theory (van Dongen 2010):

The theory of relativity is a beautiful example of the basic character of the modern development of theory. That is to say, the hypotheses from which one starts become ever more abstract and more remote from experience. But in return one comes closer to the preeminent goal of science, that of encompassing a maximum of empirical contents through logical deduction with a minimum of hypotheses or axioms. The intellectual path from the axioms to the empirical contents or to the testable consequences becomes, thereby, ever longer and more subtle. The theoretician is forced, ever more, to allow himself to be directed by purely mathematical, formal points of view in the search for theories, because the physical experience of the experimenter is not capable of leading us up to the regions of the highest abstraction. Tentative deduction takes the place of the predominantly inductive methods appropriate to the youthful state of science. Such a theoretical structure must be quite thoroughly elaborated in order for it to lead to consequences that can be compared with experience. It is certainly the case that here, as well, the empirical fact is the all-powerful judge. But its judgment can be handed down only on the basis of great and difficult intellectual effort that first bridges the wide space between the axioms and the testable consequences. The theorist must accomplish this Herculean task with the clear understanding that this effort may only be destined to prepare the way for a death sentence for his theory. One should not reproach the theorist who undertakes such a task by calling him a fantast; instead, one must allow him his fantasizing, since for him there is no other way to his goal whatsoever. Indeed, it is no planless fantasizing, but rather a search for the logically simplest possibilities and their consequences. (Einstein 1954, 238–239; Howard’s translation)

What warrant is there for thus trusting in simplicity? At best one can do a kind of meta-induction. That “the totality of all sensory experience can be ‘comprehended’ on the basis of a conceptual system built on premises of great simplicity” will be derided by skeptics as a “miracle creed,” but, Einstein adds, “it is a miracle creed which has been borne out to an amazing extent by the development of science” (Einstein 1950, p. 342). The success of previous physical theories justifies our trusting that nature is the realization of the simplest that is mathematically conceivable

But for all that Einstein’s faith in simplicity was strong, he despaired of giving a precise, formal characterization of how we assess the simplicity of a theory. In 1946 he wrote about the perspective of simplicity (here termed the “inner perfection” of a theory):

This point of view, whose exact formulation meets with great difficulties, has played an important role in the selection and evaluation of theories from time immemorial. The problem here is not simply one of a kind of enumeration of the logically independent premises (if anything like this were at all possible without ambiguity), but one of a kind of reciprocal weighing of incommensurable qualities.… I shall not attempt to excuse the lack of precision of [these] assertions … on the grounds of insufficient space at my disposal; I must confess herewith that I cannot at this point, and perhaps not at all, replace these hints by more precise definitions. I believe, however, that a sharper formulation would be possible. In any case it turns out that among the “oracles” there usually is agreement in judging the “inner perfection” of the theories and even more so concerning the degree of “external confirmation.” (Einstein 1946, pp. 21, 23).

As in 1918, so in 1946 and beyond, Einstein continues to be impressed that the “oracles,” presumably the leaders of the relevant scientific community, tend to agree in their judgments of simplicity. That is why, in practice, simplicity seems to determine theory choice univocally.

In the physics and philosophy of science literature of the late nineteenth and early twentieth centuries, the principle according to which scientific theorizing should strive for a univocal representation of nature was widely and well known under the name that it was given in the title of a widely-cited essay by Joseph Petzoldt, “The Law of Univocalness” [“Das Gesetz der Eindeutigkeit”] (Petzoldt 1895). An indication that the map of philosophical positions was drawn then in a manner very different from today is to found in the fact that this principle found favor among both anti-metaphysical logical empiricists, such as Carnap, and neo-Kantians, such as Cassirer. It played a major role in debates over the ontology of general relativity and was an important part of the background to the development of the modern concept of categoricity in formal semantics (for more on the history, influence, and demise of the principle of univocalness, see Howard 1992 and 1996). One can find no more ardent and consistent champion of the principle than Einstein.

The principle of univocalness should not be mistaken for a denial of the underdetermination thesis. The latter asserts that a multiplicity of theories can equally well account for a given body of empirical evidence, perhaps even the infinity of all possible evidence in the extreme, Quinean version of the thesis. The principle of univocalness asserts (in a somewhat anachronistic formulation) that any one theory, even any one among a set of empirically equivalent theories, should provide a univocal representation of nature by determining for itself an isomorphic set of models. The unambiguous determination of theory choice by evidence is not the same thing as the univocal determination of a class of models by a theory.

The principle of univocalness played a central role in Einstein’s struggles to formulate the general theory of relativity. When, in 1913, Einstein wrongly rejected a fully generally covariant theory of gravitation, he did so in part because he thought, wrongly, that generally covariant field equations failed the test of univocalness. More specifically, he reasoned wrongly that for a region of spacetime devoid of matter and energy—a “hole”—generally covariant field equations permit the construction of two different solutions, different in the sense that, in general, for spacetime points inside the hole, they assign different values of the metric tensor to one and the same point (for more on the history of this episode, see Stachel 1980 and Norton 1984). But Einstein’s “hole argument” is wrong, and his own diagnosis of the error in 1915 rests again, ironically, on a deployment of the principle of univocalness. What Einstein realized in 1915 was that, in 1913, he was wrongly assuming that a coordinate chart sufficed to fix the identity of spacetime manifold points. The application of a coordinate chart cannot suffice to individuate manifold points precisely because a coordinate chart is not an invariant labeling scheme, whereas univocalness in the representation of nature requires such invariance (see Howard and Norton 1993 and Howard 1999 for further discussion).

Here is how Einstein explained his change of perspective in a letter to Paul Ehrenfest of 26 December 1915, just a few weeks after the publication of the final, generally covariant formulation of the general theory of relativity:

In §12 of my work of last year, everything is correct (in the first three paragraphs) up to that which is printed with emphasis at the end of the third paragraph. From the fact that the two systems \(G(x)\) and \(G'(x)\), referred to the same reference system, satisfy the conditions of the grav. field, no contradiction follows with the univocalness of events. That which was apparently compelling in these reflections founders immediately, if one considers that the reference system signifies nothing real that the (simultaneous) realization of two different \(g\)-systems (or better, two different grav. fields) in the same region of the continuum is impossible according to the nature of the theory. In place of §12, the following reflections must appear. The physically real in the universe of events (in contrast to that which is dependent upon the choice of a reference system) consists in spatiotemporal coincidences .* [Footnote *: and in nothing else!] Real are, e.g., the intersections of two different world lines, or the statement that they do not intersect. Those statements that refer to the physically real therefore do not founder on any univocal coordinate transformation. If two systems of the \(g_{\mu v}\) (or in general the variables employed in the description of the world) are so created that one can obtain the second from the first through mere spacetime transformation, then they are completely equivalent. For they have all spatiotemporal point coincidences in common, i.e., everything that is observable. These reflections show at the same time how natural the demand for general covariance is. (CPAE, Vol. 8, Doc. 173)

Einstein’s new point of view, according to which the physically real consists exclusively in that which can be constructed on the basis of spacetime coincidences, spacetime points, for example, being regarded as intersections of world lines, is now known as the “point-coincidence argument.” Einstein might have been inspired by a paper by the young mathematician Erich Kretschmann (Howard and Norton 1993; cf. Giovanelli 2013) or possibly by a conversation with Schlick (Engler and Renn, 2017). Spacetime coincidences play this privileged ontic role because they are invariant and, thus, univocally determined. Spacetime coordinates lack such invariance, a circumstance that Einstein thereafter repeatedly formulated as the claim that space and time “thereby lose the last vestige of physical reality” (see, for example, Einstein to Ehrenfest, 5 January 1916, CPAE, Vol. 8, Doc. 180).

One telling measure of the philosophical importance of Einstein’s new perspective on the ontology of spacetime is the fact that Schlick devoted his first book, Raum und Zeit in den gegenwärtigen Physik (1917), a book for which Einstein had high praise (see Howard 1984 and 1999). But what most interested Einstein was Schlick’s discussion of the reality concept. Schlick argued that Mach was wrong to regard only the elements of sensation as real. Spacetime events, individuated invariantly as spacetime coincidences, have as much or more right to be taken as real, precisely because of the univocal manner of their determination. Einstein wholeheartedly agreed, though he ventured the above-quoted suggestion that one should distinguish the two kinds of reality—that of the elements and that of the spacetime events—on the ground that if “two different peoples” pursued physics independently of one another they were fated to agree about the elements but would almost surely produce different theoretical constructions at the level of the spacetime event ontology. Note, again, that underdetermination is not a failure of univocalness. Different though they will be, each people’s theoretical construction of an event ontology would be expected to be univocal.

Schlick, of course, went on to become the founder of the Vienna Circle, a leading figure in the development of logical empiricism, a champion of verificationism. That being so, an important question arises about Schlick’s interpretation of Einstein on the univocal determination of spacetime events as spacetime coincidences. The question is this: Do such univocal coincidences play such a privileged role because of their reality or because of their observability. Clearly the former—the reality of that which is univocally determined—is important. But are univocal spacetime coincidences real because, thanks to their invariance, they are observable? Or is their observability consequent upon their invariant reality? Einstein, himself, repeatedly stressed the observable character of spacetime coincidences, as in the 26 December 1915 letter to Ehrenfest quoted above (for additional references and a fuller discussion, see Howard 1999). [ 2 ]

Schlick, still a self-described realist in 1917, was clear about the relationship between observability and reality. He distinguished macroscopic coincidences in the field of our sense experience, to which he does accord a privileged and foundational epistemic status, from the microscopic point coincidences that define an ontology of spacetime manifold points. Mapping the former onto the latter is, for Schlick, an important part of the business of confirmation, but the reality of the spacetime manifold points is in no way consequent upon their observability. Indeed, how, strictly speaking, can one even talk of the observation of infinitesimal spacetime coincidences of the kind encountered in the intersection of two world lines? In fact, the order of implication goes the other way: Spacetime events individuated as spacetime coincidences are real because they are invariant, and such observability as they might possess is consequent upon their status as invariant bits of physical reality. For Einstein, and for Schlick in 1917, understanding the latter—physical reality—is the goal of physical theory.

As we have seen, Schlick’s Raum und Zeit in den gegenwärtigen Physik promoted a realistic interpretation of the ontology of general relativity. After reading the manuscript early in 1917, Einstein wrote to Schlick on 21 May that “the last section ‘Relations to Philosophy’ seems to me excellent” (CPAE, Vol. 8, Doc. 343), just the sort of praise one would expect from a fellow realist. Three years earlier, the Bonn mathematician, Eduard Study, had written another well-known, indeed very well-known defense of realism, Die realistische Weltansicht und die Lehre vom Raume (1914). Einstein read it in September of 1918. Much of it he liked, especially the droll style, as he said to Study in a letter of 17 September (CPAE, Vol. 8, Doc. 618). Pressed by Study to say more about the points where he disagreed, Einstein replied on 25 September in a rather surprising way:

I am supposed to explain to you my doubts? By laying stress on these it will appear that I want to pick holes in you everywhere. But things are not so bad, because I do not feel comfortable and at home in any of the “isms.” It always seems to me as though such an ism were strong only so long as it nourishes itself on the weakness of it counter-ism; but if the latter is struck dead, and it is alone on an open field, then it also turns out to be unsteady on its feet. So, away we go ! “The physical world is real.” That is supposed to be the fundamental hypothesis. What does “hypothesis” mean here? For me, a hypothesis is a statement, whose truth must be assumed for the moment, but whose meaning must be raised above all ambiguity . The above statement appears to me, however, to be, in itself, meaningless, as if one said: “The physical world is cock-a-doodle-doo.” It appears to me that the “real” is an intrinsically empty, meaningless category (pigeon hole), whose monstrous importance lies only in the fact that I can do certain things in it and not certain others. This division is, to be sure, not an arbitrary one, but instead …. I concede that the natural sciences concern the “real,” but I am still not a realist. (CPAE, Vol. 8, Doc. 624)

Lest there be any doubt that Einstein has little sympathy for the other side, he adds:

The positivist or pragmatist is strong as long as he battles against the opinion that there [are] concepts that are anchored in the “A priori.” When, in his enthusiasm, [he] forgets that all knowledge consists [in] concepts and judgments, then that is a weakness that lies not in the nature of things but in his personal disposition just as with the senseless battle against hypotheses, cf. the clear book by Duhem. In any case, the railing against atoms rests upon this weakness. Oh, how hard things are for man in this world; the path to originality leads through unreason (in the sciences), through ugliness (in the arts)-at least the path that many find passable. (CPAE, Vol. 8, Doc. 624)

What could Einstein mean by saying that he concedes that the natural sciences concern the “real,” but that he is “still not a realist” and that the “real” in the statement, “the physical world is real,” is an “intrinsically empty, meaningless category”?

The answer might be that realism, for Einstein, is not a philosophical doctrine about the interpretation of scientific theories or the semantics of theoretical terms. [ 3 ] For Einstein, realism is a physical postulate, one of a most interesting kind, as he explained on 18 March 1948 in a long note at the end of the manuscript of Max Born’s Waynflete Lectures, Natural Philosophy of Cause and Chance (1949), which Born had sent to Einstein for commentary:

I just want to explain what I mean when I say that we should try to hold on to physical reality. We are, to be sure, all of us aware of the situation regarding what will turn out to be the basic foundational concepts in physics: the point-mass or the particle is surely not among them; the field, in the Faraday - Maxwell sense, might be, but not with certainty. But that which we conceive as existing (’actual’) should somehow be localized in time and space. That is, the real in one part of space, A, should (in theory) somehow ‘exist’ independently of that which is thought of as real in another part of space, B. If a physical system stretches over the parts of space A and B, then what is present in B should somehow have an existence independent of what is present in A. What is actually present in B should thus not depend upon the type of measurement carried out in the part of space, A; it should also be independent of whether or not, after all, a measurement is made in A. If one adheres to this program, then one can hardly view the quantum-theoretical description as a complete representation of the physically real. If one attempts, nevertheless, so to view it, then one must assume that the physically real in B undergoes a sudden change because of a measurement in A. My physical instincts bristle at that suggestion. However, if one renounces the assumption that what is present in different parts of space has an independent, real existence, then I do not at all see what physics is supposed to describe. For what is thought to by a ‘system’ is, after all, just conventional, and I do not see how one is supposed to divide up the world objectively so that one can make statements about the parts. (Born 1969, 223–224; Howard’s translation)

Realism is thus the thesis of spatial separability, the claim that spatial separation is a sufficient condition for the individuation of physical systems, and its assumption is here made into almost a necessary condition for the possibility of an intelligible science of physics.

The postulate of spatial separability as that which undergirds the ontic independence and, hence, individual identities of the systems that physics describes was an important part of Einstein’s thinking about the foundations of physics since at least the time of his very first paper on the quantum hypothesis in 1905 (Einstein 1905a; for more detail on the early history of this idea in Einstein’s thinking, see Howard 1990b). But the true significance of the separability principle emerged most clearly in 1935, when (as hinted in the just-quoted remark) Einstein made it one of the central premises of his argument for the incompleteness of quantum mechanics (see Howard 1985 and 1989). It is not so clearly deployed in the published version of the Einstein, Podolsky, Rosen paper (1935), but Einstein did not write that paper and did not like the way the argument appeared there. Separability is, however, an explicit premise in all of Einstein’s later presentations of the argument for the incompleteness of quantum mechanics, both in correspondence and in print (see Howard 1985 for a detailed list of references).

In brief, the argument is this. Separability implies that spacelike separated systems have associated with them independent real states of affairs. A second postulate, locality, implies that the events in one region of spacetime cannot physically influence physical reality in a region of spacetime separated from the first by a spacelike interval. Consider now an experiment in which two systems, A and B, interact and separate, subsequent measurements on each corresponding to spacelike separated events. Separability implies that A and B have separate real physical states, and locality implies that the measurement performed on A cannot influence B’s real physical state. But quantum mechanics ascribes different theoretical states, different wave functions, to B depending upon that parameter that is measured on A. Therefore, quantum mechanics ascribes different theoretical states to B, when B possesses, in fact, one real physical state. Hence quantum mechanics is incomplete.

One wants to ask many questions. First, what notion of completeness is being invoked here? It is not deductive completeness. It is closer in kind to what is termed “categoricity” in formal semantics, a categorical theory being one whose models are all isomorphic to one another. It is closer still to the principle discussed above—and cited as a precursor of the concept of categoricity—namely, the principle of univocalness, which we found doing such important work in Einstein’s quest for a general theory of relativity, where it was the premise forcing the adoption of an invariant and thus univocal scheme for the individuation of spacetime manifold points.

The next question is why separability is viewed by Einstein as virtually an a priori necessary condition for the possibility of a science of physics. One reason is because a field theory like general relativity, which was Einstein’s model for a future unified foundation for physics, is an extreme embodiment of the principle of separability: “Field theory has carried out this principle to the extreme, in that it localizes within infinitely small (four-dimensional) space-elements the elementary things existing independently of the one another that it takes as basic, as well as the elementary laws it postulates for them” (Einstein 1948, 321–322). And a field theory like general relativity can do this because the infinitesimal metric interval—the careful way to think about separation in general relativistic spacetime—is invariant (hence univocally determined) under all continuous coordinate transformations.

Another reason why Einstein would be inclined to view separability as an a priori necessity is that, in thus invoking separability to ground individuation, Einstein places himself in a tradition of so viewing spatial separability with very strong Kantian roots (and, before Kant, Newtonian roots), a tradition in which spatial separability was known by the name that Arthur Schopenhauer famously gave to it, the principium individuationis (for a fuller discussion of this historical context, see Howard 1997).

A final question one wants to ask is: “What does any of this have to do with realism?” One might grant Einstein’s point that a real ontology requires a principle of individuation without agreeing that separability provides the only conceivable such principle. Separability together with the invariance of the infinitesimal metric interval implies that, in a general relativistic spacetime, there are joints everywhere, meaning that we can carve up the universe in any way we choose and still have ontically independent parts. But quantum entanglement can be read as implying that this libertarian scheme of individuation does not work. Can quantum mechanics not be given a realistic interpretation? Many would say, “yes.” Einstein said, “no.”

There is much that is original in Einstein’s philosophy of science as described thus far. At the very least, he rearranged the bits and pieces of doctrine that he learned from others—Kant, Mach, Duhem, Poincaré, Schlick, and others—in a strikingly novel way. But Einstein’s most original contribution to twentieth-century philosophy of science lies elsewhere, in his distinction between what he termed “principle theories” and “constructive theories.”

This idea first found its way into print in a brief 1919 article in the Times of London (Einstein 1919). A constructive theory, as the name implies, provides a constructive model for the phenomena of interest. An example would be kinetic theory. A principle theory consists of a set of individually well-confirmed, high-level empirical generalizations, “which permit of precise formulation” (Einstein 1914, 749). Examples include the first and second laws of thermodynamics. Ultimate understanding requires a constructive theory, but often, says Einstein, progress in theory is impeded by premature attempts at developing constructive theories in the absence of sufficient constraints by means of which to narrow the range of possible constructive theories. It is the function of principle theories to provide such constraint, and progress is often best achieved by focusing first on the establishment of such principles. According to Einstein, that is how he achieved his breakthrough with the theory of relativity, which, he says, is a principle theory, its two principles being the relativity principle and the light principle.

While the principle theories-constructive theories distinction first made its way into print in 1919, there is considerable evidence that it played an explicit role in Einstein’s thinking much earlier (Einstein 1907, Einstein to Sommerfeld 14 January 1908, CPAE, vol. 5, Doc. 73, Einstein 1914). Nor was it only the relativity and light principles that served Einstein as constraints in his theorizing. Thus, he explicitly mentions also the Boltzmann principle, \(S = k \log W\), as another such:

This equation connects thermodynamics with the molecular theory. It yields, as well, the statistical probabilities of the states of systems for which we are not in a position to construct a molecular-theoretical model. To that extent, Boltzmann’s magnificent idea is of significance for theoretical physics … because it provides a heuristic principle whose range extends beyond the domain of validity of molecular mechanics. (Einstein 1915, p. 262).

Einstein is here alluding the famous entropic analogy whereby, in his 1905 photon hypothesis paper, he reasoned from the fact that black body radiation in the Wien regime satisfied the Boltzmann principle to the conclusion that, in that regime, radiation behaved as if it consisted of mutually independent, corpuscle-like quanta of electromagnetic energy. The quantum hypothesis is a constructive model of radiation; the Boltzmann principle is the constraint that first suggested that model.

There are anticipations of the principle theories-constructive theories distinction in the nineteenth-century electrodynamics literature, James Clerk Maxwell, in particular, being a source from which Einstein might well have drawn (see Harman 1998). At the turn of the century, the “physics of principles” was a subject under wide discussion. At the turn of 1900, Hendrik A. Lorentz (Lorentz 1900, 1905; see Frisch 2005) and Henri Poincaré (for example, Poincaré 1904; see, Giedymin 1982, Darrigol 1995) presented the opposition between the “physics of principles“ and the “physics of models“ as commonplace. In a similar vein, Arnold Sommerfeld opposed a “physics of problems“, a style of doing physics based on concrete puzzle solving, to the “practice of principles“ defended by Max Planck (Seth 2010). Philipp Frank (1908, relying on Rey 1909) defined relativity theory as a “ conceptual theory“ based on abstract, but empirically well confirmed principles rather than on intuitive models. Probably many other examples could be find. . But however extensive his borrowings (no explicit debt was ever acknowledged), in Einstein’s hands the distinction becomes a methodological tool of impressive scope and fertility. What is puzzling, and even a bit sad, is that this most original methodological insight of Einstein’s had comparatively little impact on later philosophy of science or practice in physics. Only in recent decades, Einstein constructive-principle distinction has attracted interest in the philosophical literature, originating a still living philosophical debate on the foundation of spacetime theories (Brown 2005, Janssen 2009, Lange 2014). [ 4 ]

Einstein’s influence on twentieth-century philosophy of science is comparable to his influence on twentieth-century physics (Howard 2014). What made that possible? One explanation looks to the institutional and disciplinary history of theoretical physics and the philosophy of science. Each was, in its own domain, a new mode of thought in the latter nineteenth century, and each finally began to secure for itself a solid institutional basis in the early twentieth century. In a curious way, the two movements helped one another. Philosophers of science helped to legitimate theoretical physics by locating the significant cognitive content of science in its theories. Theoretical physicists helped to legitimate the philosophy of science by providing for analysis a subject matter that was radically reshaping our understanding of nature and the place of humankind within it. In some cases the help was even more direct, as with the work of Einstein and Max Planck in the mid-1920s to create in the physics department at the University of Berlin a chair in the philosophy of science for Reichenbach (see Hecht and Hartmann 1982). And we should remember the example of the physicists Mach and Ludwig Boltzmann who were the first two occupants of the new chair for the philosophy of science at the University of Vienna at the turn of the century.

Another explanation looks to the education of young physicists in Einstein’s day. Not only was Einstein’s own youthful reading heavily focused on philosophy, more generally, and the philosophy of science, in particular (for an overview, see Einstein 1989, xxiv–xxv; see also Howard 1994b), in which respect he was not unlike other physicists of his generation, but also his university physics curriculum included a required course on “The Theory of Scientific Thought” (see Einstein 1987, Doc. 28). An obvious question is whether or not the early cultivation of a philosophical habit of mind made a difference in the way Einstein and his contemporaries approached physics. As indicated by his November 1944 letter to Robert Thorton quoted at the beginning of this article, Einstein thought that it did.

Einstein’s letters and manuscripts, if unpublished, are cited by their numbers in the Einstein Archive (EA) control index and, if published, by volume, document number, and, if necessary, page number in:

Works by year

  • Born, Max, 1949. Natural Philosophy of Cause and Chance , Oxford: Oxford University Press.
  • Brown, Harvey R., 2005. Physical Relativity. Space-time Structure from a Dynamical Perspective , Oxford: Clarendon Press.
  • ––– (ed.), 1969. Albert Einstein-Hedwig und Max Born: Friefwechsel, 1916–1955 , Munich: Nymphenburger.
  • Canales, Jimena, 2015. Einstein, Bergson and the Debate That Changed Our Understanding of Time , Princeton: Princeton University Press.
  • Carnap, Rudolf, 1928. Der logische Aufbau der Welt , Berlin-Schlachtensee: Weltkreis-Verlag; English translation: The Logical Structure of the World & Psuedoproblems in Philosophy , Rolf A. George (trans.), Berkeley and Los Angeles: University of California Press, 1969.
  • Darrigol, Olivier, 1995. “Henri Poincaré’s Criticism of fin de siécle Electrodynamics”, Studies in History and Philosophy of Science (Part B: Studies in History and Philosophy of Modern Physics), 26 (1): 1–44.
  • Duhem, Pierre, 1906. La Théorie physique: son objet et sa structure , Paris: Chevalier & Rivière. English translation of the 2nd. ed. (1914): The Aim and Structure of Physical Theory , P. P. Wiener (trans.), Princeton, NJ: Princeton University Press, 1954; reprinted, New York: Athaneum, 1962.
  • –––, 1908. Ziel und Struktur der physikalischen Theorien , Friedrich Adler (trans.), foreword by Ernst Mach, Leipzig: Johann Ambrosius Barth.
  • Elsbach, Alfred, 1924. Kant und Einstein. Untersuchungen über das Verhältnis der modernen Erkenntnistheorie zur Relativitätstheorie , Berlin and Leipzig: Walter de Gruyter.
  • Engler, Fynn Ole and Jürgen Renn, 2013. “Hume, Einstein und Schlick über die Objektivität der Wissenschaft”, in Moritz Schlick–Die Rostocker Jahre und ihr Einfluss auf die Wiener Zeit , Fynn Ole Engler and Mathias Iven (eds.), Leipzig: Leipziger Universitätsverlag, 123–156.
  • Fine, Arthur, 1986. “Einstein’s Realism”, in The Shaky Game: Einstein, Realism, and the Quantum Theory , Chicago: University of Chicago Press, 86–111.
  • Frank, Philipp, 1909. “Die Stellung Des Relativitätsprinzips Im System Der Mechanik Und Der Elektrodynamik” Sitzungsberichte der Akademie der Wissenschaften 118 (IIa), 373–446.
  • Friedman, Michael, 1983. Foundations of Space-Time Theories: Relativistic Physics and Philosophy of Science , Princeton, NJ: Princeton University Press.
  • Frisch, Mathias, 2005. “Mechanisms, Principles, and Lorentz’s Cautious Realism”, Studies in History and Philosophy of Science (Part B: Studies in History and Philosophy of Modern Physics), 36: 659–679.
  • Giedymin, Jerzy, 1982. “The Physics of the Principles and Its Philosophy: Hamilton, Poincaré and Ramsey”, in Science and Convention: Essays on Henri Poincaré’s Philosophy of Science and the Conventionalist Tradition , Oxford: Pergamon, 42–89.
  • Giovanelli, Marco, 2013. “Erich Kretschmann as a Proto-Logical-Empiricist. Adventures and Misadventures of the Point-Coincidence Argument”, Studies in History and Philosophy of Science. Part B: Studies in History and Philosophy of Modern Physics , 44 (2), 115–134.
  • –––, 2013. “Talking at Cross-Purposes. How Einstein and the Logical Empiricists never Agreed on what they were Disagreeing About”, Synthese 190 (17): 3819–3863.
  • –––, 2014. “‘But One Must Not Legalize the Mentioned Sin’. Phenomenological vs. Dynamical Treatments of Rods and Clocks in Einstein’s Thought”, Studies in History and Philosophy of Science (Part B: Studies in History and Philosophy of Modern Physics), 48: 20–44.
  • –––, 2016. “‘…But I StillCan’t Get Rid of a Sense of Artificiality’: The Einstein-Reichenbach Debate on the Geometrization of the Electromagnetic Field”, Studies in History and Philosophy of Science. Part B: Studies in History and Philosophy of Modern Physics , 54, 35–51.
  • –––, 2018. “Physics Is a Kind of Metaphysics”, Émile Meyerson and Einstein’s late Rationalistic Realism”, European Journal for Philosophy of Science , 8: 783–829
  • Harman, P. M., 1998. The Natural Philosophy of James Clerk Maxwell , Cambridge: Cambridge University Press.
  • Hecht, Hartmut and Hoffmann, Dieter, 1982. “Die Berufung Hans Reichenbachs an die Berliner Universität”, Deutsche Zeitschrift für Philosophie 30: 651–662.
  • Holton, Gerald, 1968. “Mach, Einstein, and the Search for Reality”, Daedalus 97: 636–673. Reprinted in Thematic Origins of Scientific Thought: Kepler to Einstein , Cambridge, MA: Harvard University Press, 1973, 219–259.
  • Howard, Don, 1984. “Realism and Conventionalism in Einstein’s Philosophy of Science: The Einstein-Schlick Correspondence”, Philosophia Naturalis 21: 618–629.
  • –––, 1985. “Einstein on Locality and Separability”, Studies in History and Philosophy of Science 16: 171–201.
  • –––, 1989. “Holism, Separability, and the Metaphysical Implications of the Bell Experiments”, in Philosophical Consequences of Quantum Theory: Reflections on Bell’s Theorem , James T. Cushing and Ernan McMullin (eds.), Notre Dame, IN: University of Notre Dame Press, 224–253.
  • –––, 1990a. “Einstein and Duhem”, Synthese 83: 363–384.
  • –––, 1990b. “’Nicht sein kann was nicht sein darf,’ or the Prehistory of EPR, 1909–1935: Einstein’s Early Worries about the Quantum Mechanics of Composite Systems”, in Sixty-Two Years of Uncertainty: Historical, Philosophical, and Physical Inquiries into the Foundations of Quantum Mechanics , Proceedings of the 1989 Conference, “Ettore Majorana” Centre for Scientific Culture, International School of History of Science, Erice, Italy, 5–14 August. Arthur Miller, ed. New York: Plenum, 61–111.
  • –––, 1992. “Einstein and Eindeutigkeit: A Neglected Theme in the Philosophical Background to General Relativity”, in Jean Eisenstaedt and A. J. Kox (eds.), Studies in the History of General Relativity (Einstein Studies: Volume 3), Boston: Birkhäuser, 154–243.
  • –––, 1993. “Was Einstein Really a Realist?” Perspectives on Science: Historical, Philosophical, Social 1: 204–251.
  • –––, 1994a. “Einstein, Kant, and the Origins of Logical Empiricism”, in Language, Logic, and the Structure of Scientific Theories (Proceedings of the Carnap-Reichenbach Centennial, University of Konstanz, 21–24 May 1991), Wesley Salmon and Gereon Wolters (eds.), Pittsburgh: University of Pittsburgh Press; Konstanz: Universitätsverlag, 45–105.
  • –––, 1994b. “’A kind of vessel in which the struggle for eternal truth is played out’-Albert Einstein and the Role of Personality in Science”, in The Natural History of Paradigms: Science and the Process of Intellectual Evolution , John H. Langdon and Mary E. McGann (eds.), Indianapolis: University of Indianapolis Press, 1994, 111–138.
  • –––, 1996. “Relativity, Eindeutigkeit, and Monomorphism: Rudolf Carnap and the Development of the Categoricity Concept in Formal Semantics”, in Origins of Logical Empiricism (Minnesota Studies in the Philosophy of Science, Volume 16), Ronald N. Giere and Alan Richardson (eds.), Minneapolis and London: University of Minnesota Press, 115–164.
  • –––, 1997. “A Peek behind the Veil of Maya: Einstein, Schopenhauer, and the Historical Background of the Conception of Space as a Ground for the Individuation of Physical Systems”, in The Cosmos of Science: Essays of Exploration (Pittsburgh-Konstanz Series in the Philosophy and History of Science, Volume 6), John Earman and John D. Norton, (eds.), Pittsburgh: University of Pittsburgh Press; Konstanz: Universitätsverlag, 87–150.
  • –––, 1998. “Astride the Divided Line: Platonism, Empiricism, and Einstein’s Epistemological Opportunism”, in Idealization in Contemporary Physics (Poznan Studies in the Philosophy of the Sciences and the Humanities: Volume 63), Niall Shanks (ed.), Amsterdam and Atlanta: Rodopi, 143–163.
  • –––, 1999. “Point Coincidences and Pointer Coincidences: Einstein on Invariant Structure in Spacetime Theories”, in History of General Relativity IV: The Expanding Worlds of General Relativity (Based upon the Fourth International Conference, Berlin, Germany 31 July-3 August 1995), Hubert Goenner, Jürgen Renn, Jim Ritter, and Tilman Sauer (eds.), Boston: Birkhäuser, 463–500.
  • –––, 2014. “Einstein and the Development of Twentieth-century Philosophy of Science”, in The Cambridge Companion to Einstein , Michel Janssen and Christoph Lehner (eds.), Cambridge: Cambridge University Press, 354–376.
  • Howard, Don and Norton, John, 1993. “Out of the Labyrinth? Einstein, Hertz, and the Göttingen Answer to the Hole Argument”, in The Attraction of Gravitation. New Studies in the History of General Relativity (Einstein Studies: Volume 5), John Earman, Michel Jannsen, and John Norton (eds.),Boston: Birkhäuser, 30–62.
  • Howard, Don and Stachel, John (eds.), 1989. Einstein and the History of General Relativity (Einstein Studies: Volume 1), Boston: Birkhäuser.
  • Janssen, Michel, 2009. “Drawing the Line between Kinematics and Dynamics in Special Relativity”, Studies in History and Philosophy of Science. Part B: Studies in History and Philosophy of Modern Physics , 40 (1), 26–52.
  • Janssen, Michel and Jürgen Renn, 2007. “Untying the Knot. How Einstein Found His Way Back to Field Equations Discarded in the Zurich Notebook”, in: The Genesis of General Relativity Jürgen Renn et al. (eds.), 4 volumes, Dordrecht: Springer 839–925.
  • Lange, Marc, 2014. “Did Einstein Really Believe That Principle Theories Are Explanatorily Powerless?”, Perspectives on Science 22 (4), 449–63.
  • Lehmkuhl, Dennis, 2014. “Why Einstein Did Not Believe That General Relativity Geometrizes Gravity”, Studies in History and Philosophy of Science. Part B: Studies in History and Philosophy of Modern Physics ,, 46: 316–326.
  • Le Roy, Édouard, 1901. “Un positivisme nouveau”, Revue de Métaphysique et de Morale 9: 138–153.
  • Lorentz, Hendrik Antoon, 1900. “Electromagnetische theorieën van natuurkundige verschijnselen” Jaarboek der Rijksuniversiteit te Leiden , Bijlagen; repr. in Leiden: Brill 1900; German translation in Physikalische Zeitschrift , 1 (1900): 498–501, 514–519.
  • –––, 1905. “La thermodynamique et les théories cinétiques.“ Bulletin des séances de la Société française de physique , 35–63.
  • Mach, Ernst, 1886. Beiträge zur Analyse der Empfindungen , Jena: Gustav Fischer.
  • –––, 1897. Die Mechanik in ihrer Entwickelung historisch-kritisch dargestellt , 3rd impr. and enl. ed. Leipzig: Brockhaus.
  • –––, 1900. Die Analyse der Empfindungen und das Verhältniss des Physischen zum Psychischen , 2nd edition of Mach 1886, Jena: Gustav Fischer; English translation of the 5th edition of 1906, The Analysis of Sensations and the Relation of the Physical to the Psychical , Cora May Williams and Sydney Waterlow, trans. Chicago and London: Open Court, 1914. Reprint: New York: Dover, 1959.
  • –––, 1906. Erkenntnis und Irrtum. Skizzen zur Psychologie der Forschung , 2nd ed. Leipzig: Johann Ambrosius Barth; English translation, Knowledge and Error: Sketches on the Psychology of Enquiry , Thomas J. McCormack and Paul Foulkes, (trans.), Dordrecht and Boston: D. Reidel, 1976.
  • Meyerson, Émile, Meyerson, 1925. La déduction relativiste , Paris: Payot; Eng. tr. Meyerson 1985.
  • –––, 1985. The Relativistic Deduction.Epistemological Implications of the Theory of Relativity , Eng. tr. by David A. and Mary-Alice Sipfle, Dordrecht: Reidel.
  • Norton, John, 1984. “How Einstein Found His Field Equations”, Historical Studies in the Physical Sciences 14: 253–316. Reprinted in Howard and Stachel 1989, 101–159.
  • –––, 2000. “’Nature is the Realisation of the Simplest Conceivable Mathematical Ideas’: Einstein and the Canon of Mathematical Simplicity”, Studies in History and Philosophy of Modern Physics 31B: 135–170.
  • Petzoldt, Joseph, 1895. “Das Gesetz der Eindeutigkeit”, Vierteljahrsschrift für wissenschaftliche Philosophie und Soziologie 19: 146–203.
  • Poincaré, Henri, 1901. “Sur les Principes de la Mecanique”, Bibliotheque du Congrès Internationale de Philosophie , Sec. 3, Logique et Histoire des Sciences , Paris: A. Colin. Reprinted as: “La Mécanique classique”, in La Science et l’Hypothese , Paris: Flammarion, 1902, 110–134; English translation: “The Classical Mechanics”, n Science and Hypothesis , W. J. Greenstreet (trans.), London and New York: Walter Scott, 1905, 89–110. Reprint: New York: Dover, 1952.
  • –––, 1904. “The Principles of Mathematical Physics”, in Congress of Arts and Science, Universal Exposition, St. Louis, 1904 ( Philosophy and Mathematics : Volume 1), Howard J. Rogers, (ed.), Boston and New York: Houghton, Mifflin and Company, 1905, 604–622.
  • Quine, Willard van Orman, 1951. “Two Dogmas of Empiricism”, Philosophical Review , 60: 20–43; reprinted in From a Logical Point of View , Cambridge, MA: Harvard University Press, 1953, 20–46.
  • Reichenbach, Hans, 1920. Relativitätstheorie und Erkenntnis Apriori , Berlin: Julius Springer; English translation: The Theory of Relativity and A Priori Knowledge , Maria Reichenbach (trans. and ed.), Berkeley and Los Angeles: University of California Press, 1965.
  • –––, 1924. Axiomatik der relativistischen Raum-Zeit-Lehre ( Die Wissenschaft : Volume 72), Braunschweig: Friedrich Vieweg und Sohn; English translation: Axiomatization of the Theory of Relativity , Maria Reichenbach (trans.), Berkeley and Los Angeles: University of California Press, 1969.
  • –––, 1928. Philosophie der Raum-Zeit-Lehre , Berlin: Julius Springer; English translation, The Philosophy of Space & Time , Maria Reichenbach and John Freund (trans.), New York: Dover, 1957.
  • –––, 1949. “The Philosophical Significance of the Theory of Relativity”, in Schilpp 1949, 289–311.
  • Rey, Abel, 1907. La théorie de la physique chez les physiciens contemporains , Paris: Alcan.
  • –––, 1908. Die Theorie der Physik bei den modernen Physikern , Ger. tr. of Rey 1907, by Rudolf Eisler. Leipzig: Klinkhardt.
  • Ryckman, Thomas, 2005. The Reign of Relativity. Philosophy in Physics 1915–1925 , Oxford and New York: Oxford University Press.
  • –––, 2017. Einstein , New York: Routledge.
  • Sauer,Tilman, 2014. “Einstein’s Unified field Theory Program” in The Cambridge Companion to Einstein , Michel Janssen and Christoph Lehner (eds.), Cambridge: Cambridge University Press, 2014 281;–305.
  • Schilpp, Paul Arthur (ed.), 1949. Albert Einstein: Philosopher-Scientist (The Library of Living Philosophers: Volume 7), Evanston, IL: The Library of Living Philosophers.
  • Schlick, Moritz, 1910. “Das Wesen der Wahrheit nach der modernen Logik”, Vierteljahrsschrift für wissenschaftliche Philosophie und Soziologie 34: 386–477; English translation, “The Nature of Truth in Modern Logic”, in Schlick 1979, vol. 1, 41–103.
  • –––, 1915. “Die philosophische Bedeutung des Relativitätsprinzips”, Zeitschrift für Philosophie und philosophische Kritik 159: 129–175. English translation: “The Philosophical Significance of the Principle of Relativity”, in Schlick 1979, vol. 1, 153–189.
  • –––, 1917. Raum und Zeit in den gegenwärtigen Physik. Zur Einführung in das Verständnis der allgemeinen Relativitätstheorie , Berlin: Julius Springer; English translation of the 3rd edition, Space and Time in Contemporary Physics: An Introduction to the Theory of Relativity and Gravitation , Henry L. Brose (trans.), London and New York: Oxford University Press, 1920; reprinted in Schlick 1979, vol. 1, 207–269.
  • –––, 1921. “Kritizistische oder empiristische Deutung der neuen Physik”, Kant-Studien 26: 96–111. English translation: “Critical or Empiricist Interpretation of Modern Physics”, in Schlick 1979, vol. 1, 322–334.
  • –––, 1979. Philosophical Papers , 2 volumes, Henk L. Mulder and Barbara F. B. van de Velde-Schlick (eds.), Peter Heath (trans.), Dordrecht and Boston: D. Reidel.
  • Seth, Suman, 2010. Crafting the Quantum , Cambridge, Mass.: MIT Press.
  • Stachel, John, 1980. “Einstein’s Search for General Covariance, 1912–1915” (paper delivered at the Ninth International Conference on General Relativity and Gravitation, Jena, Germany (DDR), 17 July 1980), in Howard and Stachel 1989, 63–100.
  • Study, Eduard, 1914. Die realistische Weltansicht und die Lehre vom Raume. Geometrie, Anschauung und Erfahrung ( Die Wissenschaft : Volume 54), Braunschweig: Friedrich Vieweg & Sohn.
  • van Dongen, Jeroen, 2002. Einstein’s Unification: General Relativity and the Quest for Mathematical Naturalness , Ph.D. Dissertation, University of Amsterdam.
  • –––, 2010. Einstein’s Unification , Cambridge and New York: 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.
  • Einstein Papers Project , maintained by the California Institute of Technology.

Cassirer, Ernst | equivalence of mass and energy | general relativity: early philosophical interpretations of | geometry: in the 19th century | Natorp, Paul | physics: holism and nonseparability | quantum mechanics: Copenhagen interpretation of | quantum theory: philosophical issues in | quantum theory: the Einstein-Podolsky-Rosen argument in | space and time: absolute and relational space and motion, post-Newtonian theories | space and time: conventionality of simultaneity | space and time: inertial frames | space and time: the hole argument | Uncertainty Principle

Copyright © 2019 by Don A. Howard Marco Giovanelli < marco . giovanelli @ unito . it >

  • 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

William H. Miller III Department of Philosophy

Particles and waves: historical essays in the philosophy of science.

Particles and Waves: Historical Essays in the Philosophy of Science

  • Peter Achinstein (author)
  • Oxford University Press , 1991
  • Purchase Online

This volume brings together 11 essays by the distinguished philosopher of science, Peter Achinstein. The unifying theme is the nature of the philosophical problems surrounding the postulation of unobservable entities such as light waves, molecules, and electrons. How, if at all, is it possible to confirm scientific hypotheses about “unobservables”? Achinstein examines this question as it arose in actual scientific practice in three nineteenth-century episodes: the debate between particle and wave theorists of light, Maxwell’s kinetic theory of gases, and J.J. Thomson’s discovery of the electron. The book contains three parts, each devoted to one of these topics, beginning with an essay presenting the historical background of the episode and an introduction to the philosophical issues. There is an illuminating evaluation of various scientific methodologies, including hypothetico-deductivism, inductivism, and the method of independent warrant which combines features of the first two. Achinstein assesses the philosophical validity of both nineteenth-century and modern answers to questions about unobservables, and presents and defends his own solutions.

Science and Philosophy: A Love–Hate Relationship

  • Open access
  • Published: 02 August 2019
  • Volume 25 , pages 297–314, ( 2020 )

Cite this article

You have full access to this open access article

essay about philosophy of science

  • Sebastian De Haro   ORCID: orcid.org/0000-0002-3000-5967 1 , 2  

67k Accesses

12 Citations

49 Altmetric

Explore all metrics

In this paper I review the problematic relationship between science and philosophy; in particular, I will address the question of whether science needs philosophy, and I will offer some positive perspectives that should be helpful in developing a synergetic relationship between the two. I will review three lines of reasoning often employed in arguing that philosophy is useless for science: (a) philosophy’s death diagnosis (‘philosophy is dead’); (b) the historic-agnostic argument/challenge “show me examples where philosophy has been useful for science, for I don’t know of any”; (c) the division of property argument (or: philosophy and science have different subject matters, therefore philosophy is useless for science). These arguments will be countered with three contentions to the effect that the natural sciences need philosophy. I will: (a) point to the fallacy of anti-philosophicalism (or: ‘in order to deny the need for philosophy, one must do philosophy’) and examine the role of paradigms and presuppositions (or: why science can’t live without philosophy); (b) point out why the historical argument fails (in an example from quantum mechanics, alive and kicking); (c) briefly sketch some domains of intersection of science and philosophy and how the two can have mutual synergy. I will conclude with some implications of this synergetic relationship between science and philosophy for the liberal arts and sciences.

Similar content being viewed by others

essay about philosophy of science

Philosophy’s Shame: Reflections on an Ambivalent/Ambiviolent Relationship with Science

essay about philosophy of science

Introduction: The Armchair and the Pickaxe

Wilfrid sellars and the task of philosophy.

Avoid common mistakes on your manuscript.

1 Introduction

In this paper I will argue that: (i) The natural sciences need philosophy; and (ii) That scient ists need philosophy. I will also address some possible consequences of these theses for the Liberal Arts and Sciences. In doing so, I will have to define the sense in which I mean that science ‘needs’ philosophy and make a distinction between different ways in which different aspects or branches of science need philosophy. Most of my examples will be from physics. This being part of my professional bias, I claim that the arguments that apply to physics apply to biology, earth science, and other natural sciences as well. As I will argue, the most important distinction to be made is not between one natural science and another, but between fundamental and applied science. Once this distinction is made, the harm of treating all of the sciences en bloc, on the model of physics, can be minimized.

Why should I be defending the use of philosophy? After all, the thesis that philosophy is useful for science is not likely to be agreed upon by all practicing scientists. Science, not philosophy, is widely regarded as the more secure source of knowledge. For it has a method for declaring theories wrong: in other words, for falsifying its results. This method is called Experiment. And science has given us machines, abundant energy, technology, and a healthy attitude of scepticism. The scientific worldview has freed us from prejudice, ignorance, and the ironclad rule of authority. Natural scientists, not philosophers, have earned the trust of the public opinion in matters of truth, learning, and understanding. Experimental results, and not the scholastic distinctions of the philosophers, are the final judges in the court of Science. This, at least, would seem to be a widely held view, and partly for good reasons. So why care about philosophy after all?

The relation between science and philosophy is an intricate and somewhat problematic subject, Footnote 1 as I will review in the next Section. On the one hand, some great scientists have been great philosophers—not necessarily in the professional sense, but in the sense of deep thinking: science and philosophy often went together in the work of great figures such as Newton and Leibniz, so that it is sometimes hard—and perhaps unnecessary, and certainly anachronistic—to say where science ends and where philosophy begins. But on the other hand, philosophy is often regarded as useless, so that a philosophical outlook is irrelevant for science at best, and harmful at worst—as evinced by long pages of armchair philosophy that is blissfully uninformed by science. Or so the story goes. Hence my topic of the ‘love–hate relationship’.

Here I will concentrate on basic aspects of the relation between the two, and reduce the applications to education to a few final considerations. Reaching clarity about the fact that philosophy is useful for science is by itself an important and urgent task. Understanding this relationship is a first key step toward developing a synergetic relationship between the two fields.

2 Science Doesn’t Need Philosophy

Let us start with the objections to the first thesis above, that the sciences need philosophy. There are various good reasons why scientists can claim—and have claimed—that science does not need philosophy or that, more or less equivalently, philosophy is useless for science. I will consider three lines of reasoning here: the argument from the decline or death of philosophy, the historical or empirical argument, and the argument based on the contention that science and philosophy have different objects and methods.

The death of philosophy

Stephen Hawking has declared the official ‘death’ of philosophy, seemingly echoing Nietzsche’s famous ‘God is dead’. Commenting on questions such as the behaviour of the universe and the nature of reality, Hawking writes: “Traditionally these are questions for philosophy, but philosophy is dead. Philosophy has not kept up with modern developments in science, particularly physics. Scientists have become the bearers of the torch of discovery in our quest for knowledge.” (Hawking and Mlodinov 2010 , p. 5). In this argument, knowledge must be grounded on natural science. Questions such as, “what is the nature of physical reality”, “what things are out there in the world?” are questions that used to be within the domain of philosophy, but are now part of science. Something in philosophy must therefore be missing, without which philosophy is left a ‘dead’ discipline. And when a discipline is dead, one might just as well ignore it. Footnote 2 Since philosophers haven’t kept up with modern science, they have cut themselves off from our most secure source of knowledge and discovery. Hence the question: can we dismiss Hawking’s provocative suggestion that philosophers have largely been neglecting the natural sciences, thereby manoeuvring themselves towards a margin of irrelevance, if in our world the natural sciences are becoming increasingly dominant?

The historic-agnostic or empirical argument

The historic-agnostic argument is more cautious, and can be summarized as an agnostic stance about the usefulness of philosophy for science. It amounts to something like this: “I have never seen any examples of the usefulness of philosophy for science or, when I have seen usefulness in anything that philosophers were saying about science, it was because they were doing science not philosophy”. The argument can be appended with an enumeration of instances where the limited scope of a philosophical framework hampered progress in science and perhaps also a theoretical account of why that was the case.

Examples to this effect would seem to abound. Think of Plato’s requirement, expressed in the Timaeus , that the movements of the planets should be taken to be based on uniform circular motions. This mathematical postulate was grounded on the philosophical and theological doctrine that the most perfect motion was circular, because the motion of the mind when it reflects upon itself is circular (Plato (T) 34a, 36c, 40a). It became apparent very early on that this axiom was untenable for concentric spheres. Hipparchus and Ptolemy had to add a contrived system of eccentrics and epicycles to save the phenomena. Nevertheless, they still formally adhered to the Platonic postulate, which has been seen by many as a hampering factor in the progress of cosmology (Dijksterhuis 1950 , Part I, II D 15 and III C 68).

Another example could be Descartes’ requirement that all of physics should be based on the mechanical interactions between corpuscula with no other properties than form, size, and quantity. The dictum that physical interactions ought to take place by local contact collided with Newton’s theory of gravity, which entailed action at a distance. The dictum led Descartes to formulate his clever and imaginative, but arbitrary and unexplanatory theory of gravity on the basis of vortices, and a theory of magnetism based on the supposed screw-shapes of particles. Making the observed macroscopic phenomena supervene on microscopic details that were unobservable, and could therefore be amended at zero risk, allowed him to give qualitative and imaginative explanations of those phenomena: but he fell short of finding quantitative descriptions—let alone predictions. It took Newton to show, in Book II of the Principia, that Descartes’ vortex theory was not only physically inconsistent—additional external forces would be required to keep the vortices moving—but also inconsistent with Kepler’s laws. Richard Westfall gave the following fulminating evaluation of Descartes’s philosophy in connection to mechanics: “Most of the major steps forward in mechanics during the [17th] century involved the contradiction of Descartes. Although the mechanical philosophy asserted that the particles of matter of which the universe is composed are governed in their motions by the laws of mechanics, the precise description of motions led repeatedly to conflict between the science of mechanics and the mechanical philosophy.” (Westfall 1971 , p. 138). Again, one might take this as an instance where philosophy constrains scientific progress by its adherence to pre-conceived and non-negotiable ontological ideas.

In a third, more recent example, Lawrence Krauss has argued that, when it comes to the most philosophical questions about for instance quantum mechanics, such as ‘what is a measurement?’, he finds the reflections of physicists more useful than those of philosophers (Krauss 2012 ), again reflecting the agnostic stance that says: “Show me examples where philosophy has been useful for science, for I don’t know of any.”

The historical argument, then, generally amounts to the following: “Look at the relationship between science and philosophy in the past. Any attempts at close collaboration or integration between science and philosophy have always failed. It is useless to try.”

Division of property: method and subject matters

The third argument lies at the root of the other two. It says that philosophy and the natural sciences have different subject matters, therefore a small basis of overlap: they can live happily together without interfering with each other. This would explain the tendency of philosophers, signalled by Hawking, to retreat into the study of human affairs and human societies, leaving the study of nature to natural scientists.

The underlying reasoning can be understood as follows. The traditional distinction, at any rate since the nineteenth century, between the natural sciences and the humanities, is in their subject matters: nature would be the subject matter of the natural sciences, whereas humanities would busy themselves with the products of the human mind. The social sciences would then focus on human behaviour and social realities. Science would only be interested in brute matters of fact and not in social or linguistic constructs, and it would know those matters of fact by means of experiments carried out under certain conditions and subject to requirements of transparency and replicability. Placing philosophy in the camp of the humanities and the social sciences as opposed to the natural sciences: it also would deal with products of the mind and social constructs. This would both institutionalise philosophy’s independence from science, as well as establish its uselessness for science.

To this difference in subject matters corresponds a difference in method, emphasized by Wilhelm Dilthey: erklären (to explain) would be the task of the natural sciences; the humanities would instead aim at verstehen (to understand, or comprehend): not to give a reductive account in terms of causal relation, but to create a comprehensive view where parts can be related to the whole. Science would aim at formulating general laws of nature via the universal language of mathematics; a universal language that, even if it would include probabilistic laws, would admit of no ambiguities; the goal of the scientist would then be: to explain the behaviour of nature in terms of laws that can be falsified or verified by experiments. It should be said that this methodological argument can be held quite independently from differences in subject matters.

Philosophical interpretation of science would, according to some, either be mere speculation, reflective of our lack of knowledge, or a matter of subjectivity and personal taste: therefore irrelevant for science. In a more permissive vein along the same line of reasoning, one might concede that there are interpretational issues and matters of debate in science, but maintain that they only concern the human aspects of science, the use humans make of science: matters of ethics or subjective meaning of concepts; interpretations, being quite independent of the truth itself that science discovers, do not or should not have any significant bearing on science. Debates would result from uncertainty and lack of knowledge, rather than being part of science.

3 Biting the Bullet? Characterising ‘Philosophy’

Maybe these criticisms are not so off the mark. Maybe we should as philosophers just bite the bullet, and accept that we have managed to make ourselves irrelevant by disengaging from the latest concerns of science (a)—perhaps because we are not interested in science (c), or because we are not good at it (b). Of course, this would be an oversimplification. For there are plenty of philosophers who are interested in science, as well as scientifically well-informed. But for the next two paragraphs, I wish to entertain the thesis that maybe those criticisms are right, before I say a bit more about how I will use the word ‘philosophy’ in the rest of the paper. Scientists often lament that philosophers are ignorant of science, that they do armchair science, that they never test their theoretical conjectures, that philosophers make empirical claims that are known to be false, etc. This may have been historically true of some philosophers like for example Hegel, and perhaps it would not be hard to find current books, written by philosophers, illustrating these shortcomings. Footnote 3 So yes, maybe Hawking’s criticism—that one mischievous and catchy sentence, “philosophy is dead… Scientists have become the bearers of the torch of discovery in our quest for knowledge”—even if outrageously oversimplifying, does have bite.

But biting or not: it remains a false oversimplification. The rest of my essay will be concerned with how science needs philosophy. That should of course not make us forget the other fact—that philosophy needs science just as badly—but this essay will put that question aside.

At this point I should say a bit more about what I mean by ‘philosophy’. Defining philosophy is not an easy task; and the nature of philosophy, or of the philosophical life, has been one of the traditional philosophical questions—it was in Athens, at any rate: but, fortunately, I will not be concerned with the nature of philosophy, as such, in this essay, but rather with parts of philosophy that are close to science.

I will be especially concerned with philosophy of science, more specifically with what one could call ‘the philosophy of X’, where X is a scientific discipline such as physics, chemistry or biology. General philosophy of science is of course particularly relevant for science, since it reflects on the nature and structure of scientific theories, and on the scientific process itself. But my main argument will be about the sub-disciplines of philosophy of science concerned with specific disciplines. In this way, philosophy of science fulfils various roles. It engages critically, at various levels, with the foundations, methods, and results of the sciences. Thus it not only makes explicit what is often only mentioned implicitly by scientific theories, but it also analyses the concepts and methods of scientific theories, and engages with the interpretation of their results. Furthermore, philosophy is, as we will see in Sect.  4 , sometimes used more constructively to develop new scientific theories—what I will call science’s ‘heuristic function’.

But saying that philosophy of science is relevant to science does not mean that the importance of philosophy is limited to philosophy of science. For philosophy of science itself builds on discussions in other parts of philosophy—not only the history of philosophy, but also ethics, epistemology, and metaphysics, to name a few. Thus, although my main argument is geared towards the relevance of philosophy of science, one should not lose these broader aspects of philosophy of science of sight. Subdivisions within philosophy are drawn for practical reasons, but when analysing specific problems they can also be artificial—as we will see in Sect.  4 b), where philosophy of physics requires discussing questions that belong to epistemology and ontology. Thus metaphysics, once banned by the logical positivists in the heyday of their youthful excesses, now thrives happily in analytic philosophy in ways that would have made Carnap, and even Quine, frown. But never mind the old glories—philosophy will never obey all your commands and prohibitions, and it will use whatever tools it can get hold of.

4 Science Needs Philosophy

I now get back to the response to the anti-philosophy arguments given in Sect.  2 . What can one answer to these arguments, which seem to echo our most endearing notions and intuitions about the nature of science? Can we really deny that science and philosophy are two different worlds; that their subject matters and methods differ? Can we deny that science seeks to explain brute matters that are quite independent of human life? Can we deny the fact that unquestioned philosophical preconceptions have at times been hampering factors of scientific progress? Of course, we can’t, as I hinted at in the previous Section: philosophy corrupts the youth—I think we all need to start from that. But that’s only one side of the story, and not the most interesting part for us.

As I will argue, the doctrine that philosophy is useless for science is not only false: it is also harmful for education, society, and ultimately science itself. I will do this by advancing three arguments for the usefulness of philosophy for the natural sciences. These arguments include refutations of the misconceptions presented in Sect.  2 . They are neither wholly original nor exhaustive, but they should be a first step towards the development of a synergetic relationship between philosophy and the natural sciences.

Given the tensions between science and philosophy, vividly expressed by physicists such as Stephen Hawking and Lawrence Krauss in recent works, trying to gain some clarity in this confused subject is by itself an important and urgent task.

Why philosophy is useful (Ad 2a))

The fallacy of anti-philosophicalism

Let me start with a simple contention that responds to a small, logical, part of the previous arguments: what I have called the fallacy of anti-philosophicalism and its refutation. The refuting argument boils down to something like this: “In order to argue that one does not need philosophy, one must do philosophy.” Indeed, a convincing argument to the effect that “philosophy is useless for science” will necessarily entail the act of philosophizing. Even if ‘useful’ is a practical notion, arguing for the uselessness of discipline A for discipline B requires philosophical knowledge about B: one needs to argue that A is irrelevant to the subject matter, method, and goals of B. To declare categorically the uselessness of philosophy for science is therefore to have complete knowledge of the goals, method, and subject matter of science. But one can only argue about what those goals and subject matter should be by doing philosophy—more specifically, philosophy of science. Furthermore, we can only infer general statements about the usefulness of philosophy for science, from the study of a limited number of historical cases, by appending that study with a philosophical argument: hence by doing philosophy, in the way that historians and philosophers of science do it.

Does this debunk the argument about the death of philosophy? I submit that it does. For he who wants to insist on philosophy being useless for science must not try to rationally argue for this conviction, but must keep it as a matter of private opinion: for as soon as he starts to rationalize his view, he must start philosophizing. If there is some truth in that, as Hawking announces, philosophy is dead—and there may be some sense in which this is true—and that “Scientists have become the bearers of the torch of discovery in our quest for knowledge”, then scientists can only do so by becoming philosophers of science themselves, hence resurrecting philosophy. Hawking acknowledges this by engaging in the discipline he has declared to be ‘dead’, thereby becoming a philosopher. Indeed, training in philosophy has at least this use, that it prevents us from being bad philosophers.

But, when arguing for the usefulness of philosophy, a logical argument is not necessarily the most convincing one. For it might lead us to replace the fallacy by a more cautious statement: “philosophy is useless for science, except for one thing: to argue that philosophy has no other use for science whatever.” Nevertheless, the fact that the former statement was false and the latter sounds arbitrary and contrived, leads us to question the soundness of an approach that declares philosophy to be close to useless. It might lead us to the idea that perhaps there is some genuine value in philosophy which is useful or even necessary for science and for scientists after all. I will defend the view that philosophy is useful to scientists, and that some amount of philosophical activity is necessary in order to construct a theoretical framework for doing science.

Paradigms and presuppositions (why science can’t live without philosophy)

The necessity of philosophy for science can easily be understood from a Kuhnian perspective on how science develops. Thomas Kuhn explicates progress in science not as a linear process of theoretical formulation and experimental verification or refutation of scientific theories, but in terms of revolutions and changes of paradigm (Kuhn 1962 ). A paradigm is for Kuhn not a cookbook recipe about the mathematical laws and mechanical workings of the universe or a set of equations and technical terms and procedures. Paradigms include ways of looking at the world, practices of instrumentation, traditions of research, shared values and beliefs about which questions are considered to be scientific. Nowadays we might want to stretch this concept even further to include institutional conditions, governmental constraints and market stimuli that may be supportive of particular paradigms. Footnote 4 Scientists working in different paradigms view the world in different ways, Kuhn has emphasized. Their basic assumptions about the kinds of entities there are in the world differ, as do the kinds of primary properties they attribute to those entities. Scientists working in different paradigms may disagree, as did Einstein and Bohr, about what makes a good theory or a good explanation; or about what it means to understand a problem. In other words, there are a wide range of ontological, epistemic, and ethical presuppositions weaved into any given scientific paradigm (for some examples of this, see Sect.  4 b). If it is the case that a paradigm cannot come to birth, gain support, defeat its competitors, consolidate, and eventually die without such a set of explicit or at least tacit presuppositions, then presuppositions must be an intrinsic and necessary part of science regarded as a pursuit of truth. Such philosophical presuppositions are contributory to scientific theories, even if the theories are formally independent of them, because axioms cannot even be formulated without an agreement, taken from common and technical language, and justified within a wider paradigm, over what the terms mean and what kinds of entities they apply to; without implicit or explicit assumptions about how the terms relate to experimentally measurable quantities; without prescriptions for how the results of the theory can be verified or falsified. Paradigms also suggest meaningful goals and open questions for the theory. Thus philosophy plays a heuristic role in the discovery of new scientific theories (de Regt 2004 ): paradigms can function as guides towards the formulation of theories that describe entities of one type or another. As de Regt has cogently argued (see also the examples in the next Section), many great scientific innovators have at some point studied the works of philosophers and developed philosophical views of their own. This did not always happen very systematically, but the interest in philosophy developed by these scientists was at least above average and in turn had an important heuristic function in the formulation of new scientific theories (de Regt 2004 ).

Implicit in the heuristic role of philosophy is also an important analytic function, as I stressed in Sect.  4 . Footnote 5 One task of philosophy is to scrutinize the concepts and presuppositions of scientific theories, to analyse and lay bare what is implicit in a particular scientific paradigm. It is a philosophical task—one which is often carried out by physicists—to clarify the concepts of space, time, matter, energy, information, causality, etc. that figure in a given theory. This analysis is philosophical in so far as it makes explicit the implicit assumptions in the uses of these concepts: assumptions that scientific theories do not themselves normally state. Hence it moves beyond the point where the concepts appear as irreducible elements in the postulates of a theory. This analytic function should ultimately allow for a further step of integration, where the concepts of one science are related to the concepts of another.

The analytic function of philosophy might not only feed back into science, but become a starting point for philosophy itself: discovering what entities science assumes there to be in the world can be a useful starting point for philosophical reflection on nature. It seems key that philosophical stances on nature and science be compatible with the kinds of objects and relations that science finds. In the example given earlier: mechanistic philosophy did not admit the concept of action at a distance because the only forces allowed by the dominant philosophical paradigm were mechanical, hence the opposition to Newton’s gravity theory; whereas Kepler’s Pythagoreanism did allow for such a concept. Footnote 6

To summarize, then, some of the tasks for philosophy that we have found in relation to science:

To allow for, indeed to naturally incorporate into its own framework and build upon, the kinds of entities that science encounters in the world, and their properties and relations; Footnote 7

To scrutinize the terms and presuppositions of science, i.e. to make explicit the implicit assumptions of scientific theories: to critically analyse and clarify what the terms used by science mean, how they are articulated, and what assumptions they require, as well as how they relate to the entities that philosophy argues there to be in the world;

To discover standards for what good theories, valid modes of explanation, and appropriate scientific methods are: to offer an epistemology that does not thwart, but stimulates scientific progress;

To provide ethical guidance and discover (broad) goals for science;

To point out and articulate the interrelations between concepts that are found in different domains of the natural sciences as well as the social sciences and the humanities;

To explain how observations fit in the broader picture of the world, and to create a language where scientific results and broader human experience can complement and mutually enrich each other.

This list is neither exhaustive nor unique. Some of these general ideas will be instantiated in the two examples given in the next section.

The above points to a necessary relationship between science and philosophy. Science needs philosophy, as we have seen, in its two functions: heuristic, and analytic. Especially during changes of paradigm, philosophical debate will be part of the activities of science. None of this is to say that scientists need to be philosophers: most of them are not. So, philosophers may be drawn in at that point. But it is also not to say that professional philosophers should be doing all of the above tasks. Part of those tasks—surely 1 to 4—are often performed by scientists. Thus what I envisage here is a collaboration between scientists and philosophers. Indeed, I think we should be careful in distinguishing the disciplinary differences from the professional or individual ones. Saying that science, as a systematic theoretical and experimental study of the natural world, needs philosophy—which I have defined, in the analytic tradition, as the study of all the results of the sciences and humanities using the method of conceptual analysis—is not to say that each scientist requires philosophy. Philosophy may be merely a useful tool for scientists.

Why the historical argument fails: quantum information, alive and kicking (Ad 2b))

In this section, I give two examples where philosophical discussion has been genuinely contributory to science, along the lines discussed in 4a)ii. Before I do that, I will address the negative examples given in 2b)—examples where philosophy’s influence was rather hampering: the iron clad of mechanistic philosophy and Plato’s dictum that celestial motions should be along circles.

Working from 4a)ii we can now easily see that these examples in fact become a case in point: they illustrate the importance of philosophy for science. They make clear the need for having the right philosophical framework when doing science. If a conceptual enterprise such as philosophy were completely neutral, or indeed useless, to science, it could not be harmful to it in any important way either. But the fact is that: (A) some philosophical doctrines have been harmful for science while others have been productive; (B) it is impossible to have no philosophy at all (as I argued in 4a)); (C) the reason philosophy was harmful in some cases is because it was used in a positive way, according to its heuristic function from 4a)ii. And this heuristic function can indeed also be used positively. The correct course of action, then, is not to neglect philosophy—because, as per (A) and (B), philosophy can’t be neglected—but to embrace its presence and to use it in an intelligent and positive way, as in (C). It follows from (A) to (C) that philosophy must be relevant to science in its own specific way, even if it is only in the manner of setting necessary intellectual preconditions of freedom of mind, of trust in the power of reason and of experimental observation, etc. History shows that it is hard for scientists to free themselves from outdated philosophical modes of thought. This highlights the importance of investing in having a philosophical framework that allows for the kinds of entities that science encounters in the world. Specific tasks for philosophy are as listed in 4a)ii.

Next we will study positive historical examples where 4a)ii is at work, thereby refuting the historical argument formulated in 2b). To refute the historical argument, it suffices to show one example where philosophy has been genuinely contributory to the progress of science. The example will be interesting in so far as it also sheds light on why it was that philosophy contributed to science, thus instantiating elements of 4a)ii. There are many such examples. Kepler Footnote 8 and Sommerfeld were both inspired by Pythagorean philosophical ideas when working out their models of the harmonies of the solar system and of the atom, respectively. Let me here concentrate on another, more recent, example. It concerns the current revolution in quantum information technology. In the past ten years we have seen the first commercialization of quantum randomness: the first bank transaction built on the basis of a code encrypted not by the usual algorithms of classical cryptography (which rely on unproven mathematical assumptions such as the difficulty in factorizing large prime numbers), but based on the new field of quantum cryptography: a technique for encoding messages based on the notion of entanglement between particles at long distances. Quantum cryptography has been successfully developed and commercialized by several groups over the past twenty years or so.

As it turns out, the quantum information revolution is rooted in the efforts of scientists who saw philosophical enquiry as a necessary step in their quest for knowledge. There are two key moments in the history of quantum mechanics when physical progress crucially depended on asking the right philosophical questions. Let me take these two episodes as case studies of the question how philosophical ideas influence science, in terms of philosophy’s heuristic and analytic functions.

Einstein versus quantum mechanics

In 1927, conflicting views on quantum physics started to crystallize. Towards the end of the 5th Solvay conference in Brussels, Werner Heisenberg declared quantum mechanics to be a “closed theory, whose fundamental physical and mathematical assumptions are no longer susceptible of any modification” (Bacciagaluppi and Valentini 2006 , p. 437). In doing so, Heisenberg was voicing the shared feelings of his colleagues Niels Bohr, Wolfgang Pauli, and Paul Dirac, also present at the conference. But Einstein and Schrödinger would have none of it: the Copenhagen interpretation—as the new view of quantum mechanics came to be known—had philosophical implications that they deemed undesirable. Among those properties was the lack of determinacy in physical quantities and events. Also, Heisenberg and co. seemed to introduce a possible role for human observers in the definition of the concepts that went into science.

A few years later, in 1935, Einstein, Podolsky, and Rosen made the nature of their discomfort with quantum theory explicit in a famous article that came to be known as the EPR thought experiment. They considered pairs of correlated particles separated at long distances. The possibility to measure a property (for example, the momentum) of the first particle automatically gives information about the value of that property for the second particle, without measuring that property for the second particle, since the particles are in a state of correlation. And the possibility to measure the complementary property (for example, the position) of the first particle would as well determine the value of that quantity for the second particle. But because of the assumption that measurements done on the first particle cannot affect the properties of the second particle (after all, the particles are well-separated), the second particle must have had the values of its position and momentum determined before any measurements were done on the first particle (since, according to the formalism of quantum mechanics, a measurement of the first particle determines the value of that property for the other particle, in both cases). Since, according to standard quantum mechanics, a particle cannot simultaneously have determinate values for both its position and its momentum, this means that quantum mechanics is an incomplete theory: for it does not predict properties for the second particle that, according to the argument, it can clearly have.

The EPR argument is philosophical in the sense explained earlier, in Sect.  3 : for it analyses the foundations of quantum mechanics, trying to think clearly about the assumptions being made by standard quantum mechanics. But it also contains two substantive ontological assumptions. The first is what EPR call the ‘criterion of reality’ that if, using the formalism of quantum mechanics, one can predict with probability one the result of a measurement, then there is an element of physical reality corresponding to the physical quantity, with value equal to the predicted value of the measurement. The second assumption is what they call ‘locality’: namely, that elements of physical reality pertaining to one system cannot be affected by measurements performed on another system that is spacelike separated from the first.

Thus EPR’s quest was both physical and philosophical. In addition to these two ontological assumptions, they also impose ‘completeness’ as an epistemic desideratum that a theory should satisfy: namely, that ‘every element of the physical reality must have a counterpart in the physical theory’.

This led EPR to push the physical arguments farther than anybody had ever done before. The study of paradoxes borne out by thought experiments such as EPR has always played a major role in physics; but the resolution of such paradoxical situations almost invariably requires a philosophical stance about the principles and methods that are valued and deemed legitimate.

The EPR paper was truly philosophical in so far at it analysed and questioned the conceptual foundations of quantum theory. Especially EPR’s construal of the notion of completeness, and their criterion of reality, are explicit epistemic and ontological positions.

Does this mean that Einstein was being professional philosophers while he worked on that paper? Of course not. One should distinguish doing philosophy —something that, like I said before, can be done by both physicists and philosophers—from one’s professional label. Einstein was doing the philosophy that physics required at that point in time—and it was philosophy because he was reflecting on, and critically and constructively engaging with, the conceptual foundations of quantum theory. To do that, he needed philosophical tools. But he was of course also doing physics. So, by bringing philosophical methods into physics, he was advancing physics. I believe it is artificial, at such interdisciplinary intersections, to attempt to make too fixed a demarcation between physics and philosophy. Einstein was simply doing ground-breaking work that required methods from both fields.

Physics and the hippies

The next episode in this story of physics and philosophy took place many years later. After the publication of EPR, physicists continued to philosophize about the interpretation of quantum mechanics, but eventually the discussion died out. During the cold war, science and in particular physics gained much prestige. As class sizes grew, increasingly less time was spent on big questions and philosophical debates in the classrooms. While part of the reason for this decrease of attention on philosophical issues may have been pragmatic—philosophical discussions with large groups of students are hard to manage, and grading essay questions in exams is significantly more time consuming than computational questions—a vision was certainly at play about what education in science and technology should prepare students for. The interpretation of quantum mechanics was unlikely to prepare students who could provide societies with new gadgets or governments with new powerful weapons, whereas technical mastery of the formulas actually might. The old generation of physicists had received thorough training in the humanities—Werner Heisenberg once said “My mind was formed by studying philosophy, Plato and that sort of thing” (Buckley and Peat 1996 , p. 6) and they had indulged in philosophical musings about the meaning of it all. Now the new generation of strong-headed physicists uttered the war whoop “Shut up and calculate” and instructed their students to rally behind their utilitarian flag. Making gadgets was the new goal of physics.

The instrumentalist view of science regnant during the decades after the war is explained by Lee Smolin as follows: “When I learned physics in the 1970s, it was almost as if we were being taught to look down on people who thought about foundational problems. When we asked about the foundational issues in quantum theory, we were told that no one fully understood them but that concern with them was no longer part of science. The job was to take quantum mechanics as given and apply it to new problems. The spirit was pragmatic; “Shut up and calculate” was the mantra. People who couldn’t let go of their misgivings over the meaning of quantum theory were regarded as losers who couldn’t do the work.” (Smolin 2007 , p. 312).

But instrumentalism had to give way to other kinds of motivation for doing physics. Economic recession, budget cuts, and the decrease in the number of physics jobs made class sizes decrease again. Physicists once again had the time to think about the meaning of what they were doing. In a second, seemingly unrelated line of developments, the CIA, afraid that Americans would lag behind the Soviets, decided to fund laser physicist Harald Puthoff at Stanford University’s SRI lab in Menlo Park, California, for the study of psychic phenomena. Additional money came from NASA. Soon Puthoff would be associated with a third strand of events around the Bay Area. A dubious consortium of hippie physicists and quasi-crackpots formed an unlikely discussion group. They alternated their musings about all things quantum and the meaning of life with drinking parties and psychedelic drug use. They came to be known as the Fundamental Fysiks Group and eventually found a generous patron in self-help industry forerunner and multi-millionaire guru Werner Erhard. One goal of the hippie scientists was to use quantum mechanics for superluminal (faster-than-light) communication. This would include communication with their deceased colleagues. Needless to say, many of their arguments were misguided, but their contribution to physics was of lasting endurance. They not only put the interpretation of quantum mechanics on the research and teaching agenda; they analysed the EPR arguments and the important contributions to this discussion made by John Bell, David Bohm, and others, which had escaped the attention of scientists until then; they helped clarify the issues at stake, developed new thought experiments of their own, and raised awareness that quantum nonlocality might be useful in long-distance communication. Save the crucial (wrong) conclusion that superluminal communication was possible, several set-ups and techniques the hippies considered did not differ significantly from the ones that quantum communication uses nowadays. As David Kaiser has argued (Kaiser 2012 , p. xxiii), “The group’s efforts helped to bring sustained attention to the interpretation of quantum mechanics back into the classroom. And in a few critical instances, their work instigated major breakthroughs that—with hindsight—we may now recognize as laying crucial groundwork for quantum information science.”

Like Einstein, the Fundamental Fysiks Group worked at the intersection of physics and philosophy. They brought philosophical methods and literature to bear on problems in physics, and as such they did the kind of work that I argue physics periodically needs—regardless of who does that work, whether it is the physicists themselves or the professional philosophers.

The two examples illustrate some of the tasks of philosophy for science listed in Sect.  4 a)ii. Progress in fundamental issues such as entanglement and quantum communication stemmed from physicists’ willingness to engage in debates about ontological and epistemic issues such as the role of the observer, the completeness of the mathematical description of nature, the desiderata for a good description of nature, and so on. Progress not driven by such philosophical questions is hard to imagine in this case; the philosophical debate that actually took place acted as a positive, guiding force that pushed science further; fuelled by the posing of legitimate and relevant philosophical questions in their quest for new physics, by their being insistent on philosophical clarity and coherence rather than content with just technical mastery of the formulas, which was the trend of the day.

Synergy between science and philosophy (on objects and methods) (Ad 2c))

There are two sides to the objection regarding the difference between science and philosophy as forms of scholarship: subject matters on the one hand, methodology on the other.

I will be brief about the distinction in subject matter. Philosophy studies every subject matter that the sciences also study (recall my ‘philosophy of X’ from Sect.  3 ), but it does this with different aims and methods. The universe, possible universes other than our own, elementary particles, life, are all subjects of concern for both natural science and philosophy. Therefore, on those overlaps science and philosophy cannot be distinguished on the basis of their subject matters alone. The difference is often sought in their formal objects and methodologies: the earlier mentioned distinction between erklären and verstehen could be reframed as the statement that the natural sciences seek explanations in the modes of causal efficacy and material causation, whereas philosophy is interested in formal analysis, goals, and intentionality. This difference in methodology is often summed up by the mantra: ‘philosophy asks why-questions, science asks how-questions’. And I agree: their methods are different, and their aims (in particular, the specific interest from which they study ‘the same subject matter’) are also different. But I also submit that any such division cannot be made once and for all—the division is both vague at any point in time, as well as dynamical.

By declaring that there is such a division of intellectual activities, natural scientists and philosophers can comfortably go about their work without competing or stepping on each other’s shoes. But, as I am suggesting, the mantra is as comfortable as it is lacking in accuracy in fully reflecting the nature of the relationship between science and philosophy. Agreed: science and philosophy are in principle different forms of scholarship. For established fields of science such as classical mechanics or electromagnetism, there may be much truth in the statement that science is practically interested in how-questions, defined by the framework of the particular paradigm one is working in. But that is so only because a number of why-questions have been answered within the wider paradigm and are not being questioned any further. When paradigms are in the making, there is no clear-cut distinction between the scholar asking the how-questions and the scholar asking the why-questions. Any how-question may lead us to a why-question, and any answer to a why-question may lead us to answers to multiple how-questions. When placed in front of a why-question in the quest for a new theory, the scientist cannot retreat into the shell of specialism. He or she must struggle with the question using whichever intellectual means are available. He or she may need to establish, as the founding fathers of quantum mechanics attempted to do, what a measurement is before they can convincingly argue that there is such a thing as uncertainty in the microscopic world. The scientific quest presupposes having a number of philosophical issues settled first: or, at least, it presupposes engaging with the various conceptual options, and taking a stance on them. In so doing, the subject matters and methods of philosophers and of scientists become entangled: the relationship between science and philosophy becomes dynamical.

This is particularly true in our time, when science has expanded into realms—from far-away galaxies to the multiverse to neuroscience to molecular engineering—that were unknown territory just a number of decades before. Science is aimed at truth about the natural world, and although methodological distinctions can be made formally, one must be aware of their limitations: in particular, it would be wrong to conclude that a methodological distinction allows us to dismiss philosophy for the sake of science.

This brings us to another point: if science needs philosophy, scientific results should also be the starting point of philosophical reflection about nature. It is probably here that Hawking’s criticism of philosophy has an important core of truth to it (see footnote 1).

There is another reason why science needs philosophy. Scientific knowledge is not technical specialism cut off from the rest of human knowledge. The moment this happens would signal the forthcoming death of science. Scientific results constitute knowledge to be integrated into the broader human quest for answers about ourselves and about the universe. Philosophy helps the scientist articulate her findings in a kind of knowledge that can be shared with others, not experts in her field; it will help her discuss with other intellectuals and contribute to the general human task of getting to know the world and ourselves better.

To summarize my main argument so far: the relationship between science and philosophy may be in bad shape, and philosophy may be in bad shape, but it cannot be dead as long as we are trying to understand the universe around us. Historically, philosophy has been very influential for science, as has science been for philosophy. Any instances where philosophy had a negative effect on science in fact contribute to highlighting the importance of thinking carefully about the relation between philosophy and science. Science cannot do without philosophy because there are philosophical stances implicit in the presuppositions and goals of any scientific paradigm and in how theories are connected to reality: and it is the task of philosophy of science to critically engage with those presuppositions. Thus science needs philosophy to scrutinize those presuppositions, stances, and goals. And philosophical tools are sometimes required to make progress—as the EPR and quantum information revolution illustrate. Finally, science requires philosophy to connect its findings to the rest of human knowledge. Philosophy can act as a language connecting disciplines that are far away from each other.

Since the subject matters of science and of philosophy are partially overlapping, formal or methodological distinctions between science and philosophy only have limited ranges of applicability and certainly do not imply independence of the two disciplines. In other words, the boundaries between science and philosophy are not water-tight, nor should they be.

5 Liberal Arts and Sciences: Freeing the Mind

Having argued, at the end of the previous section, that science as such needs philosophy, I will now look at the implications of this statement for education. That is, I would like to add a few reflections about how scient ists need philosophy, and how this is to be reflected in education.

Let me start by examining what does and does not follow from what we have established so far. From the assertion that science needs philosophy in some way it does not follow that each individual scientist should be a skilled philosopher, or in fact should have any kind of developed skill in philosophy. A scientist faced with a philosophical question in the course of her research might choose to neglect it and still do a relatively good job at her research, at least for some time. Also, despite the fact that every scientist has a philosophy that is at least weaved into the presuppositions and goals of the given theory or paradigm that the scientist works in, perhaps appended with her own private reflections, it is true that science can be done for the sake of science with neglect of the philosophical presuppositions and for exclusively utilitarian goals. Obviously, utilitarian values do not offer a sustainable basis for science as a whole and for maintaining public trust in the meaningfulness of fundamental research. But for the individual scientist, they might just suffice. Furthermore, even in the case that the scientist has her own philosophical views, she is free to keep them private and not let them interfere with the research she is doing. In fact, scientists may work together on the same scientific problem while sustaining different ontological or epistemic presuppositions. Philosophy may be even less relevant for the applied scientist (although, especially for her, ethical issues will be important!). So, for all practical purposes, the individual scientist might get away with neglecting philosophy. What use, one might cynically enquire, will the laser physicist have in formal training in philosophy? Even taking the point that every scientist in fact makes use of philosophical thought of one kind or another—a set of ideas about the scientific practice, about the nature of the objects and relations that constitute her subject matter, etc.—one may still argue that it is enough for the individual scientist to work within the philosophical framework of a specific paradigm; to employ, in her daily work as a scientist, the intuitions that she internalized in the context of the specific paradigm or tradition in which she was trained. There is no need for receiving specific training in philosophical matters. Thus philosophy courses of the kind I have in mind cannot be seen as necessary prerequisites for any single scientist. But I argue that they are useful for them, and that scientists would benefit from them: and so, that science programmes ought to have such courses —again, without going into details, which would require a separate paper.

So, this suggests the following question. Shouldn’t the education of future scientists somehow reflect the connection that we have found between the sciences and philosophy? Indeed, particularly in the context of liberal arts and sciences, it is key that education reflects that connection. Science students in modern liberal arts and sciences programs should receive training in philosophy specific to their particular sciences. The kind of training I am arguing for here goes beyond general courses such as logic and philosophy of science, which are very important and are already part of some liberal arts and sciences curricula, as electives at least. It also goes beyond ethics, which is obviously an important training for scientists—although here one should go beyond the theoretical cocktail-party way in which some of these courses are taught, since their relevance often escapes the student. Perhaps such courses should be based more on actual scientific episodes and practices. But ethics is in itself a very large subject, and I have something else in mind here that more directly relates to my case studies: namely, philosophical reflection specific to each of the sciences, in fact specific to each particular science course a student takes. I mean courses such as ‘The Philosophy of X’, where X is a discipline or a collection of related disciplines (see the discussion in Sect.  3 ). And I would argue that such materials could also be part of every science course, rather than separate courses, and so are best taught by scientists. If one is intrepid, one might wish to add a course on theory construction: but I admit, this will not be easy, though it could be very beneficial at the graduate level.

Historically, it has been a goal of liberal arts and sciences education to educate the social, political, and intellectual elites. In our century, the liberal arts and sciences are often advertised in somewhat different, but related terms: ‘training the leaders of the future who can solve global problems’ is something one often hears as part of the institutional rhetoric about liberal arts and sciences. Selective admission procedures, small class settings, and emphasis on basic logical, argumentative, and rhetorical skills do confirm this vision. Clearly, some of these leaders will also be leaders in their respective scientific fields, whether in applied or in fundamental science. So, if the liberal arts and sciences aim at training the intellectual elites of the future, in particular they should be interested in the scientists who can really make a difference in research and scientists who will be the leaders of other scientists. More precisely; I will take a useful practical distinction made by Lee Smolin, even if I don’t agree with the broad-brush way Smolin applies it to string theory, nor with the details of his comparison with Kuhn’s idea of revolutionary science. The distinction goes back to Einstein, who wrote in a letter (letter to Robert A. Thornton 7 December 1944, EA pp. 61-574): “I fully agree with you about the significance and educational value of methodology as well as history and philosophy of science. So many people today—and even professional scientists—seem to me like someone who has seen thousands of trees but has never seen a forest. A knowledge of the historical and philosophical background gives that kind of independence from prejudices of his generation from which most scientists are suffering. This independence created by philosophical insight is—in my opinion—the mark of distinction between a mere artisan or specialist and a real seeker after truth.” Footnote 9 In this rich quotation, Einstein argues, as he did in many other occasions, for the significance of training in the history and philosophy of science, which gives the scientist independence of thought, which is precisely the kind of liberation of the mind that liberal arts programs also seek. Second, he calls this freedom of mind the mark of distinction between a mere specialist and a real seeker after truth. Smolin explicates this as follows. He divides scientists into seers and scientists who are master craftspeople. The seers are the ones leading the way, the ones who can see the whole forest, in Einstein’s words ‘the seekers after truth’. The master craftspeople are the ones who are very good at their particular trade, but have never seen a forest—be out of lack of interest or lack of sight. Smolin relates these two categories of scientists to the two types of science in Kuhn—normal science and revolutionary science. Footnote 10 In normal science, all the details of a given paradigm are explored and worked out. This is mainly the master craftspeople’s work. They explore the mine, excavate the tunnels, take out the valuable jewels in a mine that was found and planned by others. Revolutionary science, on the other hand, is the task of going into new territory, of doing the exploratory work required to establish radical new ideas; that is the work of the seers, the people who can think out of the existing paradigm—although never entirely—who can point out weaknesses in theories and propose new ways forward. Freedom of mind, among other things, is one of the characteristics of such scientists, and knowledge of history and philosophy contribute to that free way of thinking. If liberal arts and sciences programs advertise themselves as forming the leaders of the future, shouldn’t they be seeking to form master craftspeople as well as seekers, searchers of truth? Shouldn’t they be the breeding ground for scientists with a certain capacity of independence from prejudice and from the opinion of the majority as well as the ability to persuade others to pursue their radical ideals?

For a different perspective on this topic, see Kitchener ( 1988 ).

This is not what Hawking does, and the reason for it will become clear in the next section. He does not ignore philosophy, but engages with it.

I thank an anonymous referee for suggesting some of the above criticisms of current philosophy. For some examples, see Ladyman and Ross ( 2007 : pp. 17–24).

For the importance of tools and instrumentation, contexts, and power in different science cultures, see Galison and Stump ( 1996 ) and Galison ( 1997 ).

This ‘analytic’ function of philosophy does not strictly correlate with analytic philosophy. Both the analytic and continental traditions have of course been concerned with analysis of science and of its results.

This holds true despite the fact that Kepler was one of the initiators of mechanistic science, and that also Newton in various ways held mechanical views. He regarded his theory of gravity as a phenomenological, inductively generalized law of nature that would nevertheless require further explanation as to its causes.

See for examples the debates about the status of the wave-function in quantum mechanics: it is an important question whether the wave-function is a real entity existing in the world, or whether it merely represents the information about a system. This is a question that the formalism by itself does not answer, but nevertheless is important for how quantum mechanics is interpreted and used.

For a visualisation of Kepler’s model of the universe, see Katherine Brading’s Digital Visualization Project: https://katherinebrading.wordpress.com/news/digital-visualization-project .

Quoted by Smolin (2007) , pp. 310–311.

I take Smolin’s identification of the contrast of seers versus master crafspeople with revolutionary versus normal science to be merely a suggestive analogy. For there are important historical disanalogies too, which nevertheless do not militate against the point I am making about education.

Bacciagaluppi, G., & Valentini, A. (2006). Quantum theory at the crossroads. In Reconsidering the 1927 solvay conference . Cambridge: Cambridge University Press.

Buckley, P., & Peat, F. D. (1996). Glimpsing reality: Ideas in physics and the link to biology . Toronto: University of Toronto Press.

Google Scholar  

de Regt, H. (2004). Filosofie en natuurwetenschap: een haat-liefde verhouding”. In G. Buijs, M. Willemsen, R. van Woudenberg (red.), “Het Nut van de Wijsbegeerte” (pp. 16–23). Budel: Damon.

Dijksterhuis, E. J. (1950). De mechanisering van het wereldbeeld. Amsterdam University Press (2006); English translation: “The mechanization of the world picture. Oxford University Press (1969).

Einstein, A. (EA). The collected papers of Albert Einstein. Princeton: Princeton University Press 1986-present. https://plato.stanford.edu/entries/einstein-philscience/notes.html#1 .

Galison, P. (1997). Image and logic . Chicago: The University of Chicago Press.

Galison, P., & Stump, D. J. (1996). The disunity of science. Boundaries, contexts, and power . Palo Alto: Stanford University Press.

Hawking, S. W., & Mlodinov, L. (2010). The grand design. Bantam.

Kaiser, D. (2012). How the hippies saved physics. Norton.

Kitchener, R. F. (1988). The world view of contemporary physics. Does it need a new metaphysics? . New York, NY: SUNY Press.

Krauss, L. M. (2012). “ A universe from nothing” . Atria Books ; “ The consolation of philosophy ”, Scie ntific American (p. 243).

Kuhn, T. S. (1962). The structure of scientific revolutions (pp. 1962–1970). Chicago: The University of Chicago Press.

Ladyman, J., & Ross, D. (2007). Every thing must go . Oxford: Oxford University Press.

Book   Google Scholar  

Plato (T). “Timaeus”.

Smolin, L. (2007). The trouble with physics. The rise of string theory, the fall of a science, and what comes next”, mariner books .

Westfall, R. (1971). The construction of modern science. Mechanisms and mechanics . Cambridge: Cambridge University Press.

Download references

Acknowledgements

I thank the organizers of the conference Rethinking Liberal Education, where this paper was presented, for an inspiring conference. I also thank Dennis Dieks and Jeroen van Dongen for a long-term collaboration which has helped shape some of the ideas presented in this essay. I would also like to thank Palmyre Oomen and Rudi te Velde for discussions on these topics as well as thoughtful comments on the manuscript. I thank Jeremy Butterfield and two anonymous referees for their comments on the manuscript.

Author information

Authors and affiliations.

Trinity College, Cambridge, CB2 1TQ, UK

Sebastian De Haro

Department of History and Philosophy of Science, Free School Lane, Cambridge, CB2 3RH, UK

You can also search for this author in PubMed   Google Scholar

Corresponding author

Correspondence to Sebastian De Haro .

Additional information

Publisher's note.

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

Rights and permissions

Open Access This article is distributed under the terms of the Creative Commons Attribution 4.0 International License ( http://creativecommons.org/licenses/by/4.0/ ), which permits unrestricted use, distribution, and reproduction in any medium, provided you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made.

Reprints and permissions

About this article

De Haro, S. Science and Philosophy: A Love–Hate Relationship. Found Sci 25 , 297–314 (2020). https://doi.org/10.1007/s10699-019-09619-2

Download citation

Published : 02 August 2019

Issue Date : June 2020

DOI : https://doi.org/10.1007/s10699-019-09619-2

Share this article

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

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

Provided by the Springer Nature SharedIt content-sharing initiative

  • Philosophy of science
  • Science and philosophy
  • Liberal arts and sciences
  • Find a journal
  • Publish with us
  • Track your research

University of Notre Dame

Notre Dame Philosophical Reviews

  • Home ›
  • Reviews ›

Metaphysics and the Philosophy of Science: New Essays

Placeholder book cover

Matthew H. Slater and Zanja Yudell (eds.), Metaphysics and the Philosophy of Science: New Essays , Oxford University Press, 2017, 258 pp., $78.00 (hbk), ISBN 9780199363209.

Reviewed by Matteo Morganti, University of Rome 'Tre'

The ongoing trend of reflecting on the methodological basis of metaphysics constantly leads to new additions to the literature. Undoubtedly, the main recurring theme is the relationship, if any, between metaphysics and science; hence the prospects of a 'naturalistic' approach to metaphysical inquiry. This collection of essays edited by Matthew H. Slater and Zanja Yudell belongs to the growing series of contributions dealing with this latter topic.

As we are told on the cover flap, the volume "explores the role that metaphysics should play in our philosophizing about science." This normative endeavour, however, does not lead to a unitary, overarching view of metaphysics and the philosophy of science. Indeed, instead of this, one finds a series of largely unconnected essays that focus on different, specific aspects of the issue from a diverse and significantly heterogeneous range of standpoints. Thus, readers who have no prior perspective on naturalism, metaphysics, science and the philosophy of science are unlikely to gain one through reading this collection. However, they will no doubt benefit from the many stimulating suggestions offered at the more fine-grained level by authors coming "from the 'science end'" (p. 3). All the papers are well-written, and the case studies are dealt with in an informed and instructing way.

The collection opens with a useful introduction by Yudell. A brief overview is offered of relevant developments in the philosophy of the past century, and a useful taxonomy of approaches to metaphysics and the philosophy of science is proposed. Yudell also touches on the traditional issue of demarcation, and agrees with the widespread idea that no general criteria can be provided for sharply separating science and non-science, in this case metaphysics. Accordingly, he states that "philosophers of science . . . recognize metaphysics when they see it, and that is good enough" (p. 9). One wonders, however, if this is really so: on which basis do (or should) philosophers of science formulate their judgments? Why should they, in particular, be in charge of telling what counts as metaphysics? More importantly, it could be plausibly contended that, well-known problems with demarcation notwithstanding, in the case at hand one should at least try to say more. After all, it is probably borderline cases that are the most interesting for philosophers and, more generally, it seems hard to say anything truly interesting about the relationship between x and y without assuming some definition (debatable as it might be) of x and y that goes beyond the identification of putatively unproblematic instances. Be this as it may, let us now have a look at the essays contained in the collection -- which, as mentioned, are best evaluated separately.

Katherine Brading discusses Newton's work in the Principia , and argues that it represents a turning point in the history of metaphysics. In particular, says Brading, Newton's distinctions between absolute and relative, true and apparent, and mathematical and common time are instrumental to his project of defining a 'system of the world' on a scientific basis. And this entails that they are empirically grounded and never operate at a purely abstract level, disconnected from the actual world that scientists -- as well as laymen -- interact with. The foregoing leads Brading to regard the Principia as the earliest example of what she calls 'empiricist metaphysics.' This claim is interesting and well-argued (albeit with too many interim summaries of what has been said and is going to be said). Yet, a much more explicit account of the key concepts, and especially of the interplay between the empirical and the non-empirical, would have helped. Insofar as scientific theories are also always grounded on non-empirical assumptions\, that a 'system of the world' emerges from one's theory does not entail that the former complies with empiricist standards in an obvious way. Brading explicitly states, for instance, that "Newton stipulates that time 'flows equably'" (p. 38, italics added). While this lends support to her distinction between space and time based on the fact that "all rulers are perfect rulers . . .while. . . not all clocks are perfect clocks" (p. 36), the notion of stipulation points to an a priori element whose role in the context of Newton's methodology and philosophy of time (and hence, in the more general context of empiricist metaphysics) should be further elucidated.

Michael Strevens presents a fascinating argument against the 'wedding cake' compositional model of reality, according to which the world is essentially like a Lego building. Strevens claims that such a model is unable to handle complexity, and puts forward an alternative: so-called 'enion probability analysis' - a view whereby the focus is not on interacting individuals ('enions'), but rather on probabilities corresponding to "aggregate properties of populations" (p. 46). Enion probability analysis allegedly steers clear of problems with complexity, as the relevant probabilities do not track individual behaviours and interactions explicitly and are mutually independent, so allowing one to always stay within the boundaries of a manageable calculus.

In view of this, says Strevens, enion probability analysis should at least be regarded as a complement to traditional compositional views, and the same holds for all 'distributed ontologies' of this sort. The proposal is intriguing, but one would like to hear more about its precise metaphysical import. If probabilities, and the basic items of distributed ontologies more generally, are dependent on actual systems that 'bear' them, does this not mean that the latter, and consequently the compositional model of reality, remain 'ontologically dominant'? If they are basic, are we supposed to reify probabilities or, say, the superposition of different wave frequencies (p. 52)?

Next, Slater makes the compelling claim that metaphysics cannot simply be extracted from science, as whatever lesson we may learn from actual science is to be determined on grounds that are partly non-empirical. The thesis that species are extended composite objects qualifying as individuals, for example, appears compatible with what we know of the biological domain. However, it leads to the acceptance of objective indeterminacy in the world, a conclusion that cannot be evaluated on a purely empirical basis. Something similar happens, says Slater, in the case of the homeostatic property cluster view of species, which is no doubt explanatorily powerful but might be (and has been) rejected on the basis of the empirically underdetermined hypothesis that higher taxa are monophyletic. Slater's conclusion is that there is no simple recipe for naturalizing metaphysics, and "progressively naturalistic metaphysics [should be regarded] as a regulative ideal" (p. 76. Not surprisingly, Slater distances himself from the more radically naturalistic view presented in James Ladyman and Don Ross's well-known Every Thing Must Go (see also Ladyman's contribution to this volume). This sounds very sensible, and it is hoped that the claims and suggestions contained in this essay will be developed into a more systematic view in the future.

C. Kenneth Waters argues for the bold conclusion that reality has no general ontological structure spanning all scales. He first rejects the primacy given to physics by many authors (including Ladyman), and then looks at the philosophy of biology, and, in particular, at the history of the concept of gene. He notes that hypothetical science-oriented metaphysicians living in the 1930s would have concluded from an assessment of the best science available to them that the gene exists and is the fundamental unit of heredity (the 'classical gene'). Since we have now abandoned this view, replacing it with a much more flexible 'molecular gene' concept that plays several functions and lacks a precise fundamental structure, Waters concludes that (i) there is no unitary concept of gene to be found, corresponding to allegedly objective 'joints of nature'; and (ii) generalizing, one may conjecture that there is no general structure of reality.

Waters is certainly right that philosophers, including those doing metaphysics, should not look at science in the abstract, but rather consider it in its practice and historical development. Doing so no doubt fosters awareness of the fact that putative fundamental facts, levels, entities or what have you are hard to find, even from a science-informed perspective. On the other hand, Waters himself acknowledges that his realistic but pluralistic and anti-foundationalist stance is not the only option available. Indeed, one may insist that reality has a fundamental unity (or, perhaps, a unitary foundation), and prefer an historically grounded, anti-realist view of science to the idea that there is 'no general structure' (notice that Waters' own conclusion depends on historically contingent facts concerning current biology). Waters' preferences aside, the reader is left wondering how, lacking precise methodological indications, the choice is to be made.

Jenann Ismael attempts to outline a novel empiricist analysis of modal notions. Since modality is important for both philosophy and science, reductionist projects in the 'Humean' tradition -- aiming to reduce, say, laws of nature to regularities of a certain type, or chances to frequencies -- have an obvious appeal. However, according to Ismael, these projects are destined to fail, as the modal always outstrips the actual. To get out of this stalemate, Ismael puts forward a view whereby modality corresponds to 'guesses' about conditionals that we build out of the available data, with a view to using them as "partially prepared solutions to frequently encountered problems" (p. 120). Ismael regards this as middle-way option between reification and traditional reduction, focusing on hypothetical reasoning and decision rather than on abstract beliefs and knowledge. This is probably the most ambitious paper in the collection, and Ismael's suggestion surely deserves to be taken seriously. However, as in analogous cases, more needs to be said in order to establish whether or not one has truly identified a new theoretical option. On the one hand, the emphasis on decision and action suggests a pragmatic attitude that may recommend agnosticism with respect to metaphysics (Ismael explicitly refers to the instrumentalism of Dewey). On the other hand, Ismael's reluctance to accept the Humean approach appears to imply a more committed stance. But what should be added to the non-modal, then, if it is not objective modal facts?

Kyle Stanford critically assesses different ways in which metaphysics may be done in connection to science. What he calls 'scientific (or scientistic) metaphysics' aims to discover the fundamental structure of the world with the help of science, while remaining clearly distinct from the latter. It, says Stanford, problematically presupposes scientific realism. Thus, one should opt for either 'complementary metaphysics,' which attempts instead to include metaphysics in an integrated form of inquiry, each part of which depends on the others, or the 'metaphysics of science,' seeking to answer questions about the metaphysical commitments of our best scientific theories. According to Stanford, however, these too have clear shortcomings, as -- apparently -- they fail to add anything of value to what science alone contributes to the inquiry at hand.

Here, it must be pointed out, there is no clear supporting argument to be found, as Stanford relies more than is desirable on personal preference and intuition -- witness the number of expressions such as 'I suspect,' 'I do not see how' that populate the paper (pp. 134-5 in particular). Stanford claims that it is plausible to think that "most of the time the sort of examination undertaken in the metaphysics of science makes little contribution to this integrated naturalistic enterprise, at least of any sort that is broadly recognizable by contemporary scientists themselves" (p. 138). A legitimate position, for sure, but many questions remain. Does the 'most of the time' qualification not require one to draw further differentiations? And why should scientists have a priority in determining whether metaphysics is useful or not? It looks as though what is at work here is the frequent misconception that, if metaphysics is to be of interest at all, it has to be useful for/from the viewpoint of scientists (all of them? the majority? the most authoritative?). The present writer simply does not see any good reason for believing this to be the case.

Ladyman offers a welcome defence and elaboration of the key claims made in the abovementioned Every Thing Must Go . The defence is welcome especially because it does not employ the briskly polemical style of the book itself, too often accompanied by arguments that are not given detailed elaboration. Ladyman restates, and spells out further, the earlier claims that metaphysics has a methodology based on intuition and hypothetical reasoning, lacks a core of commonly accepted truths, and is merely based on a cost-benefit analysis of the pragmatic type, and that, as a result, metaphysics does not and cannot provide knowledge. Based on this, Ladyman goes on to remind the reader that the only acceptable metaphysics that can be read off from science is a metaphysics of structure, in particular a mixture of 'ontic structural realism' and 'rainforest realism' whereby Dennettian patterns play a central ontological role.

While this is all useful, some of the doubts that were raised when the book came out persist: what is really the role of intuition in metaphysics and in science? Can metaphysics not connect, at least indirectly, to empirical data, for example by providing the conceptual tools for interpreting scientific theories? (In this connection, it is surprising to read, in footnote 10, that "particles are not intrinsically individuated individuals": unless one is to presuppose an idiosyncratic conception of intrinsic individuation, this is far from having been established.) Also, why should unification be considered particularly important if pragmatic virtues more generally are not? And how exactly does it lead to an ontology of patterns? Relatedly, what is a structure, or pattern? (Should we believe that the Carnot cycle -- see the example on p. 153 -- is among the basic entities constituting reality?) Are these not genuinely metaphysical concepts whose clarification and assessment requires a milder characterisation of metaphysics than that recommended by Ladyman?

In a way sharing and reaffirming Ladyman's scepticism, Juha Saatsi compares explanatory considerations as they appear in science and in metaphysics. He argues that, although inference to the best explanation is ubiquitous, only in science is it complemented by experimental feedback, and only in science does it lead to something that may qualify as progress. Explanationism in metaphysics, adds Saatsi, also fails to find support in considerations of unification, understanding and Quinean indispensability -- if only because these are already problematic in the case of science. Saatsi's arguments merit further reflection but, as they stand, fail to establish more than a difference of degree between science and metaphysics. This, especially in view of the fact that - to repeat - the problem of demarcation remains open, hence reference to 'empirical feedback' cannot have a straightforward outcome.

Jim Woodward provides an entertaining dialogue with a view to illustrating the grounds for his agnosticism towards metaphysics, in particular when it comes to the philosophical analysis of causal notions. Woodward aims to defend a philosophy of science that focuses on actual practice without getting involved with things that are putatively 'fundamental' and 'deep.' His arguments are sensible, and many readers will probably find reasonable the recommendation not to seek the 'metaphysically primitive' and instead "use experimentation to establish conclusions about causal relationships independently of putative underlying details" (p. 208). However, a warning is in order: metaphysics isn't necessarily the sort of reductionist activity, aiming to translate anything and everything in terms of fundamental entities, that Woodward has in mind. It might also be understood, and is probably best understood, as the exploration of different ways the world might be, including some in which the very idea of a fundamental level or entity and/or the possibility of reduction are explicitly ruled out at the outset. In this sense, Woodward's convincing criticism of the Best System Analysis of natural laws is perhaps a bit misleading.

The closing chapter contains a reflection on metaphysical questions concerning scientific models, mathematical structures and fictional objects as they are used in science and studied by philosophers of science. Martin Thomson-Jones argues against what he calls the 'bracketing strategy,' which consists in accepting talk involving the relevant abstract objects, while at the same time resisting ontological commitment. His basic point is that the 'as-if' approach to models and the like cannot but be implemented on the basis of a specific understanding of such objects from the ontological point of view, i.e., by saying something substantial about their existence and properties. Instrumentalists about metaphysics are likely to respond that, even if one makes positive claims about the nature and existence of abstract objects, the resulting discourse could still be understood conditionally, hence realism still does not follow. Slightly differently, one may contend that the as-if approach requires specifying the nature of the relevant objects, but not whether they actually exist or not. Again, more methodological discussion is in order.

In conclusion, this collection contains significant, thought-provoking material, especially from the viewpoint of researchers approaching the topic of metaphysics in the philosophy of science with supporting background knowledge, and perhaps some pre-existing personal opinions and beliefs on the matter. It is also certainly of relevance for people interested in the specific topics that are discussed, such as the methodology of Newton's Principia , the status of biological concepts, or interventionist approaches to causation. Those looking for a more systematic, detailed and encompassing view of the interplay of metaphysics and science  will be disappointed, and probably also frustrated by the many questions that are left open and by the tension existing between at least some of the claims contained in the book. People belonging to this group had better look elsewhere -- or perhaps just wait until a compelling treatment is eventually offered.

Philosophy of Science in Nursing Essay

Nursing is a continuously evolving discipline, which changed its position from an occupation to a profession in the twentieth century. When thinking about nursing, one can imagine many different activities and ideas – all of them can pertain to this discipline equally. For example, for some people, nursing is the direct act of caring about people’s health through physical examination, diagnosis, or procedures. For others, however, nursing is a science field in which academics conduct research to improve population health. Thus, it is vital to accept nursing as a multidimensional discipline that takes knowledge from a variety of sciences of health and society.

As a profession, nursing can be regarded as mostly practical, especially if the nurse works in a clinic and cares for patients. In this way, the academic discipline of nursing also includes the process of teaching future nurses the necessary skills to provide healthcare services. Nevertheless, many scholars also add the ideas of nursing leadership, management, and science into the discipline’s view (McEwen & Wills, 2019). Finally, if one considers nursing as a science, the practice, while not disappearing, gives way to theory development and more prominent and more abstract questions.

Science and healthcare are inherently united in their use of the scientific method and the aim of discovery and evidence-seeking. Nursing, therefore, is also supported by science – evidence-based practice is one of the main pillars of this profession. However, questions such as what science is or how a scientific method can be reliable and applicable are also essential. This exploration into the deeper meaning behind science is termed the philosophy of science (Gray et al., 2017). In my opinion, this branch of philosophy is vital for nursing, as the latter unites the research behind healthcare and the human, holistic aspects of it. The process of caring for another human and trying to improve their health cannot follow a single unified route supported by limited evidence. People have different worldviews that impact their perception of the world, including their health.

Thus, the philosophy of science in nursing shows that absolute truth does not exist, and nurses have to understand that to connect to patients and to understand their personal growth as well. My nursing philosophy is based on the fact that this profession encompasses more than disease treatment. It is a holistic discipline to support and improve the wellness and health of individual patients and communities. Treatment, research, and advocacy contribute to this objective equally, creating a multifaceted approach to population health.

Philosophy impacts scientific research and also influences how nurses acquire or develop knowledge. This means that one’s beliefs or views of nursing and health can change how one interacts with information. Rega et al. (2017) explain that philosophy gives specific meaning to human life, illness, and health. By reviewing their philosophical approach to these ideas, nurses can acquire new knowledge, reject outdated or unsupported statements, or determine which knowledge path is more valuable in each particular situation.

Overall, nursing is widely understood as a combination of practice and theory. It unites actual perceived reality and the dimension of ethics that are difficult to measure or quantify. Therefore, each nurses’ philosophy influences how they interact with the field of nursing and their profession. My philosophy is based on holistic, patient-centered care that extends beyond treatment into advocacy and research.

Gray, J.R., Grove, S.K., & Sutherland, S. (2017). Burns and Grove’s the practice of nursing research: Appraisal, synthesis, and generation of evidence (8th ed.). Saunders Elsevier.

McEwen, M., & Wills, E. M. (2019). Theoretical basis for nursing (5 th ed.). Wolters Kluwer Health.

Rega, M. L., Telaretti, F., Alvaro, R., & Kangasniemi, M. (2017). Philosophical and theoretical content of the nursing discipline in academic education: A critical interpretive synthesis . Nurse Education Today , 57 , 74-81.

  • Chicago (A-D)
  • Chicago (N-B)

IvyPanda. (2022, July 8). Philosophy of Science in Nursing. https://ivypanda.com/essays/philosophy-of-science-in-nursing/

"Philosophy of Science in Nursing." IvyPanda , 8 July 2022, ivypanda.com/essays/philosophy-of-science-in-nursing/.

IvyPanda . (2022) 'Philosophy of Science in Nursing'. 8 July.

IvyPanda . 2022. "Philosophy of Science in Nursing." July 8, 2022. https://ivypanda.com/essays/philosophy-of-science-in-nursing/.

1. IvyPanda . "Philosophy of Science in Nursing." July 8, 2022. https://ivypanda.com/essays/philosophy-of-science-in-nursing/.

Bibliography

IvyPanda . "Philosophy of Science in Nursing." July 8, 2022. https://ivypanda.com/essays/philosophy-of-science-in-nursing/.

  • Principles of Holistic Nursing Care
  • Advocacy in Nursing (ANMC)
  • Caring in Nursing
  • Philosophy of Nursing and Caring Practice: Jean Watson’s Caring Model
  • The Holistic Health Promotion Model Overview
  • Jean Watson's Theory of Caring
  • Artistic Expression of Caring Concept
  • Advocacy on Home Health Nursing
  • Advocacy as an Essential Component of Nursing
  • Advocacy Activities
  • Becoming Familiar With the DNP Scholarly Project
  • Point of Care and the Peripheral Blood Draw Laboratory-Run Anticoagulant Testing
  • Theory and Practice Model of Human Caring by Jean Watson
  • Nurse Staffing Ratios: Policy Issue
  • Comfort Care of the Patient in Intensive Care

Teach philosophy of science

Affiliation.

  • 1 H. Holden Thorp Editor-in-Chief, Science journals.
  • PMID: 38603502
  • DOI: 10.1126/science.adp7153

Much is being made about the erosion of public trust in science. Surveys show a modest decline in the United States from a very high level of trust, but that is seen for other institutions as well. What is apparent from the surveys is that a better explanation of the nature of science-that it is revised as new data surface-would have a strong positive effect on public trust. Because scientists are so aware of this feature, it is often taken for granted that the public understands this too. A step toward addressing this problem would be revising undergraduate and graduate curricula to teach not just theories and techniques but the underlying philosophy of science as well.

Publication types

IMAGES

  1. The Philosophy of Science Essay Example

    essay about philosophy of science

  2. (PDF) Different Approaches to the Philosophy of Science

    essay about philosophy of science

  3. Philosophy of Science Notes

    essay about philosophy of science

  4. SCIE1000 Philosophy Essay

    essay about philosophy of science

  5. (PDF) Philosophy of Science : Final Examination Paper

    essay about philosophy of science

  6. Philosophy of Science

    essay about philosophy of science

VIDEO

  1. Method of philosophy vs science

  2. Essay on blessings of science for class 7-12|importance of science|Advantages of science

  3. How to Write an AMAZING Teaching Philosophy Statement || How to Write a Pedagogical Statement

  4. What is Philosophy? The Love of Wisdom

  5. Why Psychology Isn't a Science

  6. Science vs. Philosophy

COMMENTS

  1. The philosophy of science

    The philosophy of science is a field that deals with what science is, how it works, and the logic through which we build scientific knowledge. In this website, we present a rough synthesis of some new and some old ideas from the philosophy of science. The philosophy of science. Photo credit: Wikimedia. In this website, we use a practical ...

  2. Philosophy of Science, Essay Example

    Quantum mechanics is an example of objective truth where the philosophy of mathematics and physics can be put to work to test a hypothetical theory. The idea that science is objective and always changing does mean it has some fit in this world. Scientist still do put rockets in the air that work. Most of science is built on objective theory but ...

  3. Free Philosophy of Science Essay Examples & Topics

    Philosophy of science is a field that studies scientific inquiries from a philosophical perspective. To write an essay about it, you should get acquainted with some of the big names within the area. Aristotle is, possibly, the founder of the field. He wrote a lot on physics, psychology, biology, logic, and mathematics.

  4. Philosophy of science

    Philosophy of science is the branch of philosophy concerned with the foundations, methods, and implications of science. Amongst its central questions are the difference between science and non-science, the reliability of scientific theories, and the ultimate purpose and meaning of science as a human endeavour.

  5. philosophy of science

    philosophy of science, the study, from a philosophical perspective, of the elements of scientific inquiry. This article discusses metaphysical, epistemological, and ethical issues related to the practice and goals of modern science. For treatment of philosophical issues raised by the problems and concepts of specific sciences, see biology ...

  6. Thomas Kuhn

    Thomas Kuhn. Thomas Samuel Kuhn (1922-1996) is one of the most influential philosophers of science of the twentieth century, perhaps the most influential. His 1962 book The Structure of Scientific Revolutions is one of the most cited academic books of all time. Kuhn's contribution to the philosophy of science marked not only a break with ...

  7. Scientific Discovery

    Scientific discovery is the process or product of successful scientific inquiry. Objects of discovery can be things, events, processes, causes, and properties as well as theories and hypotheses and their features (their explanatory power, for example). Most philosophical discussions of scientific discoveries focus on the generation of new ...

  8. Philosophy

    Biologists balk at any talk of 'goals' or 'intentions' - but a bold new research agenda has put agency back on the table. Philip Ball. Philosophy Essays from Aeon. World-leading thinkers explore life's big questions and the history of ideas from Socrates to Simone de Beauvoir, political philosophy to philosophy of mind, the Western ...

  9. Einstein's Philosophy of Science

    Albert Einstein (1879-1955) is well known as the most prominent physicist of the twentieth century. His contributions to twentieth-century philosophy of science, though of comparable importance, are less well known. Einstein's own philosophy of science is an original synthesis of elements drawn from sources as diverse as neo-Kantianism ...

  10. PDF INTRODUCTION TO THE PHILOSOPHY OF SCIENCE

    10 STARTING OUT SOME METAPHYSICAL PREMISES OF SCIENCE v Ontological realism - the belief that reality and its components exist independently of any consciousness. (Ontology is the branch of philosophy that attempts to answer questions regarding the existence/non-existence of things and their nature). v Epistemological realism - the belief in the 'knowability' of things, which ...

  11. Philosophy of Science

    NEW TO CAMBRIDGE IN 2022 Since its inception in 1934, Philosophy of Science, along with its sponsoring society, the Philosophy of Science Association, has been dedicated to the furthering of studies and free discussion from diverse standpoints in the philosophy of science. The journal contains essays, discussion articles, and book reviews.

  12. Evidence, Explanation, and Realism: Essays in Philosophy of Science

    The essays in this volume address three fundamental questions in the philosophy of science: What is required for some fact to be evidence for a scientific hypothesis? What does it mean to say that a scientist or a theory explains a phenomenon? Should scientific theories that postulate "unobservable" entities such as electrons be construed realistically...

  13. Particles and Waves: Historical Essays in the Philosophy of Science

    This volume brings together 11 essays by the distinguished philosopher of science, Peter Achinstein. The unifying theme is the nature of the philosophical problems surrounding the postulation of unobservable entities such as light waves, molecules, and electrons. How, if at all, is it possible to confirm scientific hypotheses about "unobservables"? Achinstein examines this question as...

  14. Philosophy of Science: Paradigm, Ontology, Epistemology Essay

    Philosophy of Science: Paradigm, Ontology, Epistemology Essay. The term Philosophy of Science deals with all the foundations, methodologies, implications, and assumptions of science during a given process. On several occasions, it determines the magnitude of truth in a particular set of scientific results, thereby the merits or demerits of the ...

  15. Philosophy of Science

    Read the latest issue. Since its inception in 1934, Philosophy of Science, along with its sponsoring society, The Philosophy of Science Association, has been dedicated to the furthering of studies and free discussion from diverse standpoints in the philosophy of science. The journal contains essays, discussion articles, and book reviews.

  16. Philosophy of Science Essay Topics

    Philosophy of Science Essay Topics. Instructor Joanna Harris. Joanna has taught high school social studies both online and in a traditional classroom since 2009, and has a doctorate in Educational ...

  17. Science and Philosophy: A Love-Hate Relationship

    Defining philosophy is not an easy task; and the nature of philosophy, or of the philosophical life, has been one of the traditional philosophical questions—it was in Athens, at any rate: but, fortunately, I will not be concerned with the nature of philosophy, as such, in this essay, but rather with parts of philosophy that are close to science.

  18. Philosophy of Science Essay

    The Impact Of Science And Philosophy Of Science Essay. Alan Francis Chalmers is an associate professor who works extensive in the history and philosophy of science (physical). Alan Chalmers has taught at the University of Sydney since 1971, first in the School of Philosophy, and from 1987 at the Unit for the History and Philosophy of Science ...

  19. The Philosophy Of Science Philosophy Essay

    Philosophy of science is one of the formal philosophy that wondered about science and its scientific methods. It can be defined as the study of understanding of scientific things and its scientific processes or methods. Philosophy does not account its processes to limited knowledges, but is concerned with the variety of principles and laws of ...

  20. Metaphysics and the Philosophy of Science: New Essays

    A brief overview is offered of relevant developments in the philosophy of the past century, and a useful taxonomy of approaches to metaphysics and the philosophy of science is proposed. Yudell also touches on the traditional issue of demarcation, and agrees with the widespread idea that no general criteria can be provided for sharply separating ...

  21. The Philosophy Of Science

    The Philosophy Of Science. This essay sample was donated by a student to help the academic community. Papers provided by EduBirdie writers usually outdo students' samples. 'Describe and understand how science works within a wide range of sciences. This does not have to include every kind of science.

  22. Essay about the Philosophy of Science : r/PhilosophyofScience

    Most of the pointers given to us i mentioned in the first paragraph of the post i made. The grading criteria includes 1) show an understandinf of history of science, demarcation and the scientific method 2) reference atleast 3 of the philosophers of science (or streams). 3)reflect on the role of science in society (role of ethics, responsible ...

  23. Philosophy of Science in Nursing

    Philosophy impacts scientific research and also influences how nurses acquire or develop knowledge. This means that one's beliefs or views of nursing and health can change how one interacts with information. Rega et al. (2017) explain that philosophy gives specific meaning to human life, illness, and health.

  24. Teach philosophy of science

    Much is being made about the erosion of public trust in science. Surveys show a modest decline in the United States from a very high level of trust, but that is seen for other institutions as well. What is apparent from the surveys is that a better explanation of the nature of science-that it is rev …

  25. Philosophy of computer science on prediction of human liver disease

    The end result is axiological philosophy of science from the analysis of the truth of the performance of this study which states that the performance of the neural network has a greater accuracy of 94.4% and is followed by other methods such as nave Bayes 81.3%, kNN 93.6%, and SVM 94.1%.