What is the Scientific Method: How does it work and why is it important?

The scientific method is a systematic process involving steps like defining questions, forming hypotheses, conducting experiments, and analyzing data. It minimizes biases and enables replicable research, leading to groundbreaking discoveries like Einstein's theory of relativity, penicillin, and the structure of DNA. This ongoing approach promotes reason, evidence, and the pursuit of truth in science.

Updated on November 18, 2023

What is the Scientific Method: How does it work and why is it important?

Beginning in elementary school, we are exposed to the scientific method and taught how to put it into practice. As a tool for learning, it prepares children to think logically and use reasoning when seeking answers to questions.

Rather than jumping to conclusions, the scientific method gives us a recipe for exploring the world through observation and trial and error. We use it regularly, sometimes knowingly in academics or research, and sometimes subconsciously in our daily lives.

In this article we will refresh our memories on the particulars of the scientific method, discussing where it comes from, which elements comprise it, and how it is put into practice. Then, we will consider the importance of the scientific method, who uses it and under what circumstances.

What is the scientific method?

The scientific method is a dynamic process that involves objectively investigating questions through observation and experimentation . Applicable to all scientific disciplines, this systematic approach to answering questions is more accurately described as a flexible set of principles than as a fixed series of steps.

The following representations of the scientific method illustrate how it can be both condensed into broad categories and also expanded to reveal more and more details of the process. These graphics capture the adaptability that makes this concept universally valuable as it is relevant and accessible not only across age groups and educational levels but also within various contexts.

a graph of the scientific method

Steps in the scientific method

While the scientific method is versatile in form and function, it encompasses a collection of principles that create a logical progression to the process of problem solving:

  • Define a question : Constructing a clear and precise problem statement that identifies the main question or goal of the investigation is the first step. The wording must lend itself to experimentation by posing a question that is both testable and measurable.
  • Gather information and resources : Researching the topic in question to find out what is already known and what types of related questions others are asking is the next step in this process. This background information is vital to gaining a full understanding of the subject and in determining the best design for experiments. 
  • Form a hypothesis : Composing a concise statement that identifies specific variables and potential results, which can then be tested, is a crucial step that must be completed before any experimentation. An imperfection in the composition of a hypothesis can result in weaknesses to the entire design of an experiment.
  • Perform the experiments : Testing the hypothesis by performing replicable experiments and collecting resultant data is another fundamental step of the scientific method. By controlling some elements of an experiment while purposely manipulating others, cause and effect relationships are established.
  • Analyze the data : Interpreting the experimental process and results by recognizing trends in the data is a necessary step for comprehending its meaning and supporting the conclusions. Drawing inferences through this systematic process lends substantive evidence for either supporting or rejecting the hypothesis.
  • Report the results : Sharing the outcomes of an experiment, through an essay, presentation, graphic, or journal article, is often regarded as a final step in this process. Detailing the project's design, methods, and results not only promotes transparency and replicability but also adds to the body of knowledge for future research.
  • Retest the hypothesis : Repeating experiments to see if a hypothesis holds up in all cases is a step that is manifested through varying scenarios. Sometimes a researcher immediately checks their own work or replicates it at a future time, or another researcher will repeat the experiments to further test the hypothesis.

a chart of the scientific method

Where did the scientific method come from?

Oftentimes, ancient peoples attempted to answer questions about the unknown by:

  • Making simple observations
  • Discussing the possibilities with others deemed worthy of a debate
  • Drawing conclusions based on dominant opinions and preexisting beliefs

For example, take Greek and Roman mythology. Myths were used to explain everything from the seasons and stars to the sun and death itself.

However, as societies began to grow through advancements in agriculture and language, ancient civilizations like Egypt and Babylonia shifted to a more rational analysis for understanding the natural world. They increasingly employed empirical methods of observation and experimentation that would one day evolve into the scientific method . 

In the 4th century, Aristotle, considered the Father of Science by many, suggested these elements , which closely resemble the contemporary scientific method, as part of his approach for conducting science:

  • Study what others have written about the subject.
  • Look for the general consensus about the subject.
  • Perform a systematic study of everything even partially related to the topic.

a pyramid of the scientific method

By continuing to emphasize systematic observation and controlled experiments, scholars such as Al-Kindi and Ibn al-Haytham helped expand this concept throughout the Islamic Golden Age . 

In his 1620 treatise, Novum Organum , Sir Francis Bacon codified the scientific method, arguing not only that hypotheses must be tested through experiments but also that the results must be replicated to establish a truth. Coming at the height of the Scientific Revolution, this text made the scientific method accessible to European thinkers like Galileo and Isaac Newton who then put the method into practice.

As science modernized in the 19th century, the scientific method became more formalized, leading to significant breakthroughs in fields such as evolution and germ theory. Today, it continues to evolve, underpinning scientific progress in diverse areas like quantum mechanics, genetics, and artificial intelligence.

Why is the scientific method important?

The history of the scientific method illustrates how the concept developed out of a need to find objective answers to scientific questions by overcoming biases based on fear, religion, power, and cultural norms. This still holds true today.

By implementing this standardized approach to conducting experiments, the impacts of researchers’ personal opinions and preconceived notions are minimized. The organized manner of the scientific method prevents these and other mistakes while promoting the replicability and transparency necessary for solid scientific research.

The importance of the scientific method is best observed through its successes, for example: 

  • “ Albert Einstein stands out among modern physicists as the scientist who not only formulated a theory of revolutionary significance but also had the genius to reflect in a conscious and technical way on the scientific method he was using.” Devising a hypothesis based on the prevailing understanding of Newtonian physics eventually led Einstein to devise the theory of general relativity .
  • Howard Florey “Perhaps the most useful lesson which has come out of the work on penicillin has been the demonstration that success in this field depends on the development and coordinated use of technical methods.” After discovering a mold that prevented the growth of Staphylococcus bacteria, Dr. Alexander Flemimg designed experiments to identify and reproduce it in the lab, thus leading to the development of penicillin .
  • James D. Watson “Every time you understand something, religion becomes less likely. Only with the discovery of the double helix and the ensuing genetic revolution have we had grounds for thinking that the powers held traditionally to be the exclusive property of the gods might one day be ours. . . .” By using wire models to conceive a structure for DNA, Watson and Crick crafted a hypothesis for testing combinations of amino acids, X-ray diffraction images, and the current research in atomic physics, resulting in the discovery of DNA’s double helix structure .

Final thoughts

As the cases exemplify, the scientific method is never truly completed, but rather started and restarted. It gave these researchers a structured process that was easily replicated, modified, and built upon. 

While the scientific method may “end” in one context, it never literally ends. When a hypothesis, design, methods, and experiments are revisited, the scientific method simply picks up where it left off. Each time a researcher builds upon previous knowledge, the scientific method is restored with the pieces of past efforts.

By guiding researchers towards objective results based on transparency and reproducibility, the scientific method acts as a defense against bias, superstition, and preconceived notions. As we embrace the scientific method's enduring principles, we ensure that our quest for knowledge remains firmly rooted in reason, evidence, and the pursuit of truth.

The AJE Team

The AJE Team

See our "Privacy Policy"

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

Scientific Method

Science is an enormously successful human enterprise. The study of scientific method is the attempt to discern the activities by which that success is achieved. Among the activities often identified as characteristic of science are systematic observation and experimentation, inductive and deductive reasoning, and the formation and testing of hypotheses and theories. How these are carried out in detail can vary greatly, but characteristics like these have been looked to as a way of demarcating scientific activity from non-science, where only enterprises which employ some canonical form of scientific method or methods should be considered science (see also the entry on science and pseudo-science ). Others have questioned whether there is anything like a fixed toolkit of methods which is common across science and only science. Some reject privileging one view of method as part of rejecting broader views about the nature of science, such as naturalism (Dupré 2004); some reject any restriction in principle (pluralism).

Scientific method should be distinguished from the aims and products of science, such as knowledge, predictions, or control. Methods are the means by which those goals are achieved. Scientific method should also be distinguished from meta-methodology, which includes the values and justifications behind a particular characterization of scientific method (i.e., a methodology) — values such as objectivity, reproducibility, simplicity, or past successes. Methodological rules are proposed to govern method and it is a meta-methodological question whether methods obeying those rules satisfy given values. Finally, method is distinct, to some degree, from the detailed and contextual practices through which methods are implemented. The latter might range over: specific laboratory techniques; mathematical formalisms or other specialized languages used in descriptions and reasoning; technological or other material means; ways of communicating and sharing results, whether with other scientists or with the public at large; or the conventions, habits, enforced customs, and institutional controls over how and what science is carried out.

While it is important to recognize these distinctions, their boundaries are fuzzy. Hence, accounts of method cannot be entirely divorced from their methodological and meta-methodological motivations or justifications, Moreover, each aspect plays a crucial role in identifying methods. Disputes about method have therefore played out at the detail, rule, and meta-rule levels. Changes in beliefs about the certainty or fallibility of scientific knowledge, for instance (which is a meta-methodological consideration of what we can hope for methods to deliver), have meant different emphases on deductive and inductive reasoning, or on the relative importance attached to reasoning over observation (i.e., differences over particular methods.) Beliefs about the role of science in society will affect the place one gives to values in scientific method.

The issue which has shaped debates over scientific method the most in the last half century is the question of how pluralist do we need to be about method? Unificationists continue to hold out for one method essential to science; nihilism is a form of radical pluralism, which considers the effectiveness of any methodological prescription to be so context sensitive as to render it not explanatory on its own. Some middle degree of pluralism regarding the methods embodied in scientific practice seems appropriate. But the details of scientific practice vary with time and place, from institution to institution, across scientists and their subjects of investigation. How significant are the variations for understanding science and its success? How much can method be abstracted from practice? This entry describes some of the attempts to characterize scientific method or methods, as well as arguments for a more context-sensitive approach to methods embedded in actual scientific practices.

1. Overview and organizing themes

2. historical review: aristotle to mill, 3.1 logical constructionism and operationalism, 3.2. h-d as a logic of confirmation, 3.3. popper and falsificationism, 3.4 meta-methodology and the end of method, 4. statistical methods for hypothesis testing, 5.1 creative and exploratory practices.

  • 5.2 Computer methods and the ‘new ways’ of doing science

6.1 “The scientific method” in science education and as seen by scientists

6.2 privileged methods and ‘gold standards’, 6.3 scientific method in the court room, 6.4 deviating practices, 7. conclusion, other internet resources, related entries.

This entry could have been given the title Scientific Methods and gone on to fill volumes, or it could have been extremely short, consisting of a brief summary rejection of the idea that there is any such thing as a unique Scientific Method at all. Both unhappy prospects are due to the fact that scientific activity varies so much across disciplines, times, places, and scientists that any account which manages to unify it all will either consist of overwhelming descriptive detail, or trivial generalizations.

The choice of scope for the present entry is more optimistic, taking a cue from the recent movement in philosophy of science toward a greater attention to practice: to what scientists actually do. This “turn to practice” can be seen as the latest form of studies of methods in science, insofar as it represents an attempt at understanding scientific activity, but through accounts that are neither meant to be universal and unified, nor singular and narrowly descriptive. To some extent, different scientists at different times and places can be said to be using the same method even though, in practice, the details are different.

Whether the context in which methods are carried out is relevant, or to what extent, will depend largely on what one takes the aims of science to be and what one’s own aims are. For most of the history of scientific methodology the assumption has been that the most important output of science is knowledge and so the aim of methodology should be to discover those methods by which scientific knowledge is generated.

Science was seen to embody the most successful form of reasoning (but which form?) to the most certain knowledge claims (but how certain?) on the basis of systematically collected evidence (but what counts as evidence, and should the evidence of the senses take precedence, or rational insight?) Section 2 surveys some of the history, pointing to two major themes. One theme is seeking the right balance between observation and reasoning (and the attendant forms of reasoning which employ them); the other is how certain scientific knowledge is or can be.

Section 3 turns to 20 th century debates on scientific method. In the second half of the 20 th century the epistemic privilege of science faced several challenges and many philosophers of science abandoned the reconstruction of the logic of scientific method. Views changed significantly regarding which functions of science ought to be captured and why. For some, the success of science was better identified with social or cultural features. Historical and sociological turns in the philosophy of science were made, with a demand that greater attention be paid to the non-epistemic aspects of science, such as sociological, institutional, material, and political factors. Even outside of those movements there was an increased specialization in the philosophy of science, with more and more focus on specific fields within science. The combined upshot was very few philosophers arguing any longer for a grand unified methodology of science. Sections 3 and 4 surveys the main positions on scientific method in 20 th century philosophy of science, focusing on where they differ in their preference for confirmation or falsification or for waiving the idea of a special scientific method altogether.

In recent decades, attention has primarily been paid to scientific activities traditionally falling under the rubric of method, such as experimental design and general laboratory practice, the use of statistics, the construction and use of models and diagrams, interdisciplinary collaboration, and science communication. Sections 4–6 attempt to construct a map of the current domains of the study of methods in science.

As these sections illustrate, the question of method is still central to the discourse about science. Scientific method remains a topic for education, for science policy, and for scientists. It arises in the public domain where the demarcation or status of science is at issue. Some philosophers have recently returned, therefore, to the question of what it is that makes science a unique cultural product. This entry will close with some of these recent attempts at discerning and encapsulating the activities by which scientific knowledge is achieved.

Attempting a history of scientific method compounds the vast scope of the topic. This section briefly surveys the background to modern methodological debates. What can be called the classical view goes back to antiquity, and represents a point of departure for later divergences. [ 1 ]

We begin with a point made by Laudan (1968) in his historical survey of scientific method:

Perhaps the most serious inhibition to the emergence of the history of theories of scientific method as a respectable area of study has been the tendency to conflate it with the general history of epistemology, thereby assuming that the narrative categories and classificatory pigeon-holes applied to the latter are also basic to the former. (1968: 5)

To see knowledge about the natural world as falling under knowledge more generally is an understandable conflation. Histories of theories of method would naturally employ the same narrative categories and classificatory pigeon holes. An important theme of the history of epistemology, for example, is the unification of knowledge, a theme reflected in the question of the unification of method in science. Those who have identified differences in kinds of knowledge have often likewise identified different methods for achieving that kind of knowledge (see the entry on the unity of science ).

Different views on what is known, how it is known, and what can be known are connected. Plato distinguished the realms of things into the visible and the intelligible ( The Republic , 510a, in Cooper 1997). Only the latter, the Forms, could be objects of knowledge. The intelligible truths could be known with the certainty of geometry and deductive reasoning. What could be observed of the material world, however, was by definition imperfect and deceptive, not ideal. The Platonic way of knowledge therefore emphasized reasoning as a method, downplaying the importance of observation. Aristotle disagreed, locating the Forms in the natural world as the fundamental principles to be discovered through the inquiry into nature ( Metaphysics Z , in Barnes 1984).

Aristotle is recognized as giving the earliest systematic treatise on the nature of scientific inquiry in the western tradition, one which embraced observation and reasoning about the natural world. In the Prior and Posterior Analytics , Aristotle reflects first on the aims and then the methods of inquiry into nature. A number of features can be found which are still considered by most to be essential to science. For Aristotle, empiricism, careful observation (but passive observation, not controlled experiment), is the starting point. The aim is not merely recording of facts, though. For Aristotle, science ( epistêmê ) is a body of properly arranged knowledge or learning—the empirical facts, but also their ordering and display are of crucial importance. The aims of discovery, ordering, and display of facts partly determine the methods required of successful scientific inquiry. Also determinant is the nature of the knowledge being sought, and the explanatory causes proper to that kind of knowledge (see the discussion of the four causes in the entry on Aristotle on causality ).

In addition to careful observation, then, scientific method requires a logic as a system of reasoning for properly arranging, but also inferring beyond, what is known by observation. Methods of reasoning may include induction, prediction, or analogy, among others. Aristotle’s system (along with his catalogue of fallacious reasoning) was collected under the title the Organon . This title would be echoed in later works on scientific reasoning, such as Novum Organon by Francis Bacon, and Novum Organon Restorum by William Whewell (see below). In Aristotle’s Organon reasoning is divided primarily into two forms, a rough division which persists into modern times. The division, known most commonly today as deductive versus inductive method, appears in other eras and methodologies as analysis/​synthesis, non-ampliative/​ampliative, or even confirmation/​verification. The basic idea is there are two “directions” to proceed in our methods of inquiry: one away from what is observed, to the more fundamental, general, and encompassing principles; the other, from the fundamental and general to instances or implications of principles.

The basic aim and method of inquiry identified here can be seen as a theme running throughout the next two millennia of reflection on the correct way to seek after knowledge: carefully observe nature and then seek rules or principles which explain or predict its operation. The Aristotelian corpus provided the framework for a commentary tradition on scientific method independent of science itself (cosmos versus physics.) During the medieval period, figures such as Albertus Magnus (1206–1280), Thomas Aquinas (1225–1274), Robert Grosseteste (1175–1253), Roger Bacon (1214/1220–1292), William of Ockham (1287–1347), Andreas Vesalius (1514–1546), Giacomo Zabarella (1533–1589) all worked to clarify the kind of knowledge obtainable by observation and induction, the source of justification of induction, and best rules for its application. [ 2 ] Many of their contributions we now think of as essential to science (see also Laudan 1968). As Aristotle and Plato had employed a framework of reasoning either “to the forms” or “away from the forms”, medieval thinkers employed directions away from the phenomena or back to the phenomena. In analysis, a phenomena was examined to discover its basic explanatory principles; in synthesis, explanations of a phenomena were constructed from first principles.

During the Scientific Revolution these various strands of argument, experiment, and reason were forged into a dominant epistemic authority. The 16 th –18 th centuries were a period of not only dramatic advance in knowledge about the operation of the natural world—advances in mechanical, medical, biological, political, economic explanations—but also of self-awareness of the revolutionary changes taking place, and intense reflection on the source and legitimation of the method by which the advances were made. The struggle to establish the new authority included methodological moves. The Book of Nature, according to the metaphor of Galileo Galilei (1564–1642) or Francis Bacon (1561–1626), was written in the language of mathematics, of geometry and number. This motivated an emphasis on mathematical description and mechanical explanation as important aspects of scientific method. Through figures such as Henry More and Ralph Cudworth, a neo-Platonic emphasis on the importance of metaphysical reflection on nature behind appearances, particularly regarding the spiritual as a complement to the purely mechanical, remained an important methodological thread of the Scientific Revolution (see the entries on Cambridge platonists ; Boyle ; Henry More ; Galileo ).

In Novum Organum (1620), Bacon was critical of the Aristotelian method for leaping from particulars to universals too quickly. The syllogistic form of reasoning readily mixed those two types of propositions. Bacon aimed at the invention of new arts, principles, and directions. His method would be grounded in methodical collection of observations, coupled with correction of our senses (and particularly, directions for the avoidance of the Idols, as he called them, kinds of systematic errors to which naïve observers are prone.) The community of scientists could then climb, by a careful, gradual and unbroken ascent, to reliable general claims.

Bacon’s method has been criticized as impractical and too inflexible for the practicing scientist. Whewell would later criticize Bacon in his System of Logic for paying too little attention to the practices of scientists. It is hard to find convincing examples of Bacon’s method being put in to practice in the history of science, but there are a few who have been held up as real examples of 16 th century scientific, inductive method, even if not in the rigid Baconian mold: figures such as Robert Boyle (1627–1691) and William Harvey (1578–1657) (see the entry on Bacon ).

It is to Isaac Newton (1642–1727), however, that historians of science and methodologists have paid greatest attention. Given the enormous success of his Principia Mathematica and Opticks , this is understandable. The study of Newton’s method has had two main thrusts: the implicit method of the experiments and reasoning presented in the Opticks, and the explicit methodological rules given as the Rules for Philosophising (the Regulae) in Book III of the Principia . [ 3 ] Newton’s law of gravitation, the linchpin of his new cosmology, broke with explanatory conventions of natural philosophy, first for apparently proposing action at a distance, but more generally for not providing “true”, physical causes. The argument for his System of the World ( Principia , Book III) was based on phenomena, not reasoned first principles. This was viewed (mainly on the continent) as insufficient for proper natural philosophy. The Regulae counter this objection, re-defining the aims of natural philosophy by re-defining the method natural philosophers should follow. (See the entry on Newton’s philosophy .)

To his list of methodological prescriptions should be added Newton’s famous phrase “ hypotheses non fingo ” (commonly translated as “I frame no hypotheses”.) The scientist was not to invent systems but infer explanations from observations, as Bacon had advocated. This would come to be known as inductivism. In the century after Newton, significant clarifications of the Newtonian method were made. Colin Maclaurin (1698–1746), for instance, reconstructed the essential structure of the method as having complementary analysis and synthesis phases, one proceeding away from the phenomena in generalization, the other from the general propositions to derive explanations of new phenomena. Denis Diderot (1713–1784) and editors of the Encyclopédie did much to consolidate and popularize Newtonianism, as did Francesco Algarotti (1721–1764). The emphasis was often the same, as much on the character of the scientist as on their process, a character which is still commonly assumed. The scientist is humble in the face of nature, not beholden to dogma, obeys only his eyes, and follows the truth wherever it leads. It was certainly Voltaire (1694–1778) and du Chatelet (1706–1749) who were most influential in propagating the latter vision of the scientist and their craft, with Newton as hero. Scientific method became a revolutionary force of the Enlightenment. (See also the entries on Newton , Leibniz , Descartes , Boyle , Hume , enlightenment , as well as Shank 2008 for a historical overview.)

Not all 18 th century reflections on scientific method were so celebratory. Famous also are George Berkeley’s (1685–1753) attack on the mathematics of the new science, as well as the over-emphasis of Newtonians on observation; and David Hume’s (1711–1776) undermining of the warrant offered for scientific claims by inductive justification (see the entries on: George Berkeley ; David Hume ; Hume’s Newtonianism and Anti-Newtonianism ). Hume’s problem of induction motivated Immanuel Kant (1724–1804) to seek new foundations for empirical method, though as an epistemic reconstruction, not as any set of practical guidelines for scientists. Both Hume and Kant influenced the methodological reflections of the next century, such as the debate between Mill and Whewell over the certainty of inductive inferences in science.

The debate between John Stuart Mill (1806–1873) and William Whewell (1794–1866) has become the canonical methodological debate of the 19 th century. Although often characterized as a debate between inductivism and hypothetico-deductivism, the role of the two methods on each side is actually more complex. On the hypothetico-deductive account, scientists work to come up with hypotheses from which true observational consequences can be deduced—hence, hypothetico-deductive. Because Whewell emphasizes both hypotheses and deduction in his account of method, he can be seen as a convenient foil to the inductivism of Mill. However, equally if not more important to Whewell’s portrayal of scientific method is what he calls the “fundamental antithesis”. Knowledge is a product of the objective (what we see in the world around us) and subjective (the contributions of our mind to how we perceive and understand what we experience, which he called the Fundamental Ideas). Both elements are essential according to Whewell, and he was therefore critical of Kant for too much focus on the subjective, and John Locke (1632–1704) and Mill for too much focus on the senses. Whewell’s fundamental ideas can be discipline relative. An idea can be fundamental even if it is necessary for knowledge only within a given scientific discipline (e.g., chemical affinity for chemistry). This distinguishes fundamental ideas from the forms and categories of intuition of Kant. (See the entry on Whewell .)

Clarifying fundamental ideas would therefore be an essential part of scientific method and scientific progress. Whewell called this process “Discoverer’s Induction”. It was induction, following Bacon or Newton, but Whewell sought to revive Bacon’s account by emphasising the role of ideas in the clear and careful formulation of inductive hypotheses. Whewell’s induction is not merely the collecting of objective facts. The subjective plays a role through what Whewell calls the Colligation of Facts, a creative act of the scientist, the invention of a theory. A theory is then confirmed by testing, where more facts are brought under the theory, called the Consilience of Inductions. Whewell felt that this was the method by which the true laws of nature could be discovered: clarification of fundamental concepts, clever invention of explanations, and careful testing. Mill, in his critique of Whewell, and others who have cast Whewell as a fore-runner of the hypothetico-deductivist view, seem to have under-estimated the importance of this discovery phase in Whewell’s understanding of method (Snyder 1997a,b, 1999). Down-playing the discovery phase would come to characterize methodology of the early 20 th century (see section 3 ).

Mill, in his System of Logic , put forward a narrower view of induction as the essence of scientific method. For Mill, induction is the search first for regularities among events. Among those regularities, some will continue to hold for further observations, eventually gaining the status of laws. One can also look for regularities among the laws discovered in a domain, i.e., for a law of laws. Which “law law” will hold is time and discipline dependent and open to revision. One example is the Law of Universal Causation, and Mill put forward specific methods for identifying causes—now commonly known as Mill’s methods. These five methods look for circumstances which are common among the phenomena of interest, those which are absent when the phenomena are, or those for which both vary together. Mill’s methods are still seen as capturing basic intuitions about experimental methods for finding the relevant explanatory factors ( System of Logic (1843), see Mill entry). The methods advocated by Whewell and Mill, in the end, look similar. Both involve inductive generalization to covering laws. They differ dramatically, however, with respect to the necessity of the knowledge arrived at; that is, at the meta-methodological level (see the entries on Whewell and Mill entries).

3. Logic of method and critical responses

The quantum and relativistic revolutions in physics in the early 20 th century had a profound effect on methodology. Conceptual foundations of both theories were taken to show the defeasibility of even the most seemingly secure intuitions about space, time and bodies. Certainty of knowledge about the natural world was therefore recognized as unattainable. Instead a renewed empiricism was sought which rendered science fallible but still rationally justifiable.

Analyses of the reasoning of scientists emerged, according to which the aspects of scientific method which were of primary importance were the means of testing and confirming of theories. A distinction in methodology was made between the contexts of discovery and justification. The distinction could be used as a wedge between the particularities of where and how theories or hypotheses are arrived at, on the one hand, and the underlying reasoning scientists use (whether or not they are aware of it) when assessing theories and judging their adequacy on the basis of the available evidence. By and large, for most of the 20 th century, philosophy of science focused on the second context, although philosophers differed on whether to focus on confirmation or refutation as well as on the many details of how confirmation or refutation could or could not be brought about. By the mid-20 th century these attempts at defining the method of justification and the context distinction itself came under pressure. During the same period, philosophy of science developed rapidly, and from section 4 this entry will therefore shift from a primarily historical treatment of the scientific method towards a primarily thematic one.

Advances in logic and probability held out promise of the possibility of elaborate reconstructions of scientific theories and empirical method, the best example being Rudolf Carnap’s The Logical Structure of the World (1928). Carnap attempted to show that a scientific theory could be reconstructed as a formal axiomatic system—that is, a logic. That system could refer to the world because some of its basic sentences could be interpreted as observations or operations which one could perform to test them. The rest of the theoretical system, including sentences using theoretical or unobservable terms (like electron or force) would then either be meaningful because they could be reduced to observations, or they had purely logical meanings (called analytic, like mathematical identities). This has been referred to as the verifiability criterion of meaning. According to the criterion, any statement not either analytic or verifiable was strictly meaningless. Although the view was endorsed by Carnap in 1928, he would later come to see it as too restrictive (Carnap 1956). Another familiar version of this idea is operationalism of Percy William Bridgman. In The Logic of Modern Physics (1927) Bridgman asserted that every physical concept could be defined in terms of the operations one would perform to verify the application of that concept. Making good on the operationalisation of a concept even as simple as length, however, can easily become enormously complex (for measuring very small lengths, for instance) or impractical (measuring large distances like light years.)

Carl Hempel’s (1950, 1951) criticisms of the verifiability criterion of meaning had enormous influence. He pointed out that universal generalizations, such as most scientific laws, were not strictly meaningful on the criterion. Verifiability and operationalism both seemed too restrictive to capture standard scientific aims and practice. The tenuous connection between these reconstructions and actual scientific practice was criticized in another way. In both approaches, scientific methods are instead recast in methodological roles. Measurements, for example, were looked to as ways of giving meanings to terms. The aim of the philosopher of science was not to understand the methods per se , but to use them to reconstruct theories, their meanings, and their relation to the world. When scientists perform these operations, however, they will not report that they are doing them to give meaning to terms in a formal axiomatic system. This disconnect between methodology and the details of actual scientific practice would seem to violate the empiricism the Logical Positivists and Bridgman were committed to. The view that methodology should correspond to practice (to some extent) has been called historicism, or intuitionism. We turn to these criticisms and responses in section 3.4 . [ 4 ]

Positivism also had to contend with the recognition that a purely inductivist approach, along the lines of Bacon-Newton-Mill, was untenable. There was no pure observation, for starters. All observation was theory laden. Theory is required to make any observation, therefore not all theory can be derived from observation alone. (See the entry on theory and observation in science .) Even granting an observational basis, Hume had already pointed out that one could not deductively justify inductive conclusions without begging the question by presuming the success of the inductive method. Likewise, positivist attempts at analyzing how a generalization can be confirmed by observations of its instances were subject to a number of criticisms. Goodman (1965) and Hempel (1965) both point to paradoxes inherent in standard accounts of confirmation. Recent attempts at explaining how observations can serve to confirm a scientific theory are discussed in section 4 below.

The standard starting point for a non-inductive analysis of the logic of confirmation is known as the Hypothetico-Deductive (H-D) method. In its simplest form, a sentence of a theory which expresses some hypothesis is confirmed by its true consequences. As noted in section 2 , this method had been advanced by Whewell in the 19 th century, as well as Nicod (1924) and others in the 20 th century. Often, Hempel’s (1966) description of the H-D method, illustrated by the case of Semmelweiss’ inferential procedures in establishing the cause of childbed fever, has been presented as a key account of H-D as well as a foil for criticism of the H-D account of confirmation (see, for example, Lipton’s (2004) discussion of inference to the best explanation; also the entry on confirmation ). Hempel described Semmelsweiss’ procedure as examining various hypotheses explaining the cause of childbed fever. Some hypotheses conflicted with observable facts and could be rejected as false immediately. Others needed to be tested experimentally by deducing which observable events should follow if the hypothesis were true (what Hempel called the test implications of the hypothesis), then conducting an experiment and observing whether or not the test implications occurred. If the experiment showed the test implication to be false, the hypothesis could be rejected. If the experiment showed the test implications to be true, however, this did not prove the hypothesis true. The confirmation of a test implication does not verify a hypothesis, though Hempel did allow that “it provides at least some support, some corroboration or confirmation for it” (Hempel 1966: 8). The degree of this support then depends on the quantity, variety and precision of the supporting evidence.

Another approach that took off from the difficulties with inductive inference was Karl Popper’s critical rationalism or falsificationism (Popper 1959, 1963). Falsification is deductive and similar to H-D in that it involves scientists deducing observational consequences from the hypothesis under test. For Popper, however, the important point was not the degree of confirmation that successful prediction offered to a hypothesis. The crucial thing was the logical asymmetry between confirmation, based on inductive inference, and falsification, which can be based on a deductive inference. (This simple opposition was later questioned, by Lakatos, among others. See the entry on historicist theories of scientific rationality. )

Popper stressed that, regardless of the amount of confirming evidence, we can never be certain that a hypothesis is true without committing the fallacy of affirming the consequent. Instead, Popper introduced the notion of corroboration as a measure for how well a theory or hypothesis has survived previous testing—but without implying that this is also a measure for the probability that it is true.

Popper was also motivated by his doubts about the scientific status of theories like the Marxist theory of history or psycho-analysis, and so wanted to demarcate between science and pseudo-science. Popper saw this as an importantly different distinction than demarcating science from metaphysics. The latter demarcation was the primary concern of many logical empiricists. Popper used the idea of falsification to draw a line instead between pseudo and proper science. Science was science because its method involved subjecting theories to rigorous tests which offered a high probability of failing and thus refuting the theory.

A commitment to the risk of failure was important. Avoiding falsification could be done all too easily. If a consequence of a theory is inconsistent with observations, an exception can be added by introducing auxiliary hypotheses designed explicitly to save the theory, so-called ad hoc modifications. This Popper saw done in pseudo-science where ad hoc theories appeared capable of explaining anything in their field of application. In contrast, science is risky. If observations showed the predictions from a theory to be wrong, the theory would be refuted. Hence, scientific hypotheses must be falsifiable. Not only must there exist some possible observation statement which could falsify the hypothesis or theory, were it observed, (Popper called these the hypothesis’ potential falsifiers) it is crucial to the Popperian scientific method that such falsifications be sincerely attempted on a regular basis.

The more potential falsifiers of a hypothesis, the more falsifiable it would be, and the more the hypothesis claimed. Conversely, hypotheses without falsifiers claimed very little or nothing at all. Originally, Popper thought that this meant the introduction of ad hoc hypotheses only to save a theory should not be countenanced as good scientific method. These would undermine the falsifiabililty of a theory. However, Popper later came to recognize that the introduction of modifications (immunizations, he called them) was often an important part of scientific development. Responding to surprising or apparently falsifying observations often generated important new scientific insights. Popper’s own example was the observed motion of Uranus which originally did not agree with Newtonian predictions. The ad hoc hypothesis of an outer planet explained the disagreement and led to further falsifiable predictions. Popper sought to reconcile the view by blurring the distinction between falsifiable and not falsifiable, and speaking instead of degrees of testability (Popper 1985: 41f.).

From the 1960s on, sustained meta-methodological criticism emerged that drove philosophical focus away from scientific method. A brief look at those criticisms follows, with recommendations for further reading at the end of the entry.

Thomas Kuhn’s The Structure of Scientific Revolutions (1962) begins with a well-known shot across the bow for philosophers of science:

History, if viewed as a repository for more than anecdote or chronology, could produce a decisive transformation in the image of science by which we are now possessed. (1962: 1)

The image Kuhn thought needed transforming was the a-historical, rational reconstruction sought by many of the Logical Positivists, though Carnap and other positivists were actually quite sympathetic to Kuhn’s views. (See the entry on the Vienna Circle .) Kuhn shares with other of his contemporaries, such as Feyerabend and Lakatos, a commitment to a more empirical approach to philosophy of science. Namely, the history of science provides important data, and necessary checks, for philosophy of science, including any theory of scientific method.

The history of science reveals, according to Kuhn, that scientific development occurs in alternating phases. During normal science, the members of the scientific community adhere to the paradigm in place. Their commitment to the paradigm means a commitment to the puzzles to be solved and the acceptable ways of solving them. Confidence in the paradigm remains so long as steady progress is made in solving the shared puzzles. Method in this normal phase operates within a disciplinary matrix (Kuhn’s later concept of a paradigm) which includes standards for problem solving, and defines the range of problems to which the method should be applied. An important part of a disciplinary matrix is the set of values which provide the norms and aims for scientific method. The main values that Kuhn identifies are prediction, problem solving, simplicity, consistency, and plausibility.

An important by-product of normal science is the accumulation of puzzles which cannot be solved with resources of the current paradigm. Once accumulation of these anomalies has reached some critical mass, it can trigger a communal shift to a new paradigm and a new phase of normal science. Importantly, the values that provide the norms and aims for scientific method may have transformed in the meantime. Method may therefore be relative to discipline, time or place

Feyerabend also identified the aims of science as progress, but argued that any methodological prescription would only stifle that progress (Feyerabend 1988). His arguments are grounded in re-examining accepted “myths” about the history of science. Heroes of science, like Galileo, are shown to be just as reliant on rhetoric and persuasion as they are on reason and demonstration. Others, like Aristotle, are shown to be far more reasonable and far-reaching in their outlooks then they are given credit for. As a consequence, the only rule that could provide what he took to be sufficient freedom was the vacuous “anything goes”. More generally, even the methodological restriction that science is the best way to pursue knowledge, and to increase knowledge, is too restrictive. Feyerabend suggested instead that science might, in fact, be a threat to a free society, because it and its myth had become so dominant (Feyerabend 1978).

An even more fundamental kind of criticism was offered by several sociologists of science from the 1970s onwards who rejected the methodology of providing philosophical accounts for the rational development of science and sociological accounts of the irrational mistakes. Instead, they adhered to a symmetry thesis on which any causal explanation of how scientific knowledge is established needs to be symmetrical in explaining truth and falsity, rationality and irrationality, success and mistakes, by the same causal factors (see, e.g., Barnes and Bloor 1982, Bloor 1991). Movements in the Sociology of Science, like the Strong Programme, or in the social dimensions and causes of knowledge more generally led to extended and close examination of detailed case studies in contemporary science and its history. (See the entries on the social dimensions of scientific knowledge and social epistemology .) Well-known examinations by Latour and Woolgar (1979/1986), Knorr-Cetina (1981), Pickering (1984), Shapin and Schaffer (1985) seem to bear out that it was social ideologies (on a macro-scale) or individual interactions and circumstances (on a micro-scale) which were the primary causal factors in determining which beliefs gained the status of scientific knowledge. As they saw it therefore, explanatory appeals to scientific method were not empirically grounded.

A late, and largely unexpected, criticism of scientific method came from within science itself. Beginning in the early 2000s, a number of scientists attempting to replicate the results of published experiments could not do so. There may be close conceptual connection between reproducibility and method. For example, if reproducibility means that the same scientific methods ought to produce the same result, and all scientific results ought to be reproducible, then whatever it takes to reproduce a scientific result ought to be called scientific method. Space limits us to the observation that, insofar as reproducibility is a desired outcome of proper scientific method, it is not strictly a part of scientific method. (See the entry on reproducibility of scientific results .)

By the close of the 20 th century the search for the scientific method was flagging. Nola and Sankey (2000b) could introduce their volume on method by remarking that “For some, the whole idea of a theory of scientific method is yester-year’s debate …”.

Despite the many difficulties that philosophers encountered in trying to providing a clear methodology of conformation (or refutation), still important progress has been made on understanding how observation can provide evidence for a given theory. Work in statistics has been crucial for understanding how theories can be tested empirically, and in recent decades a huge literature has developed that attempts to recast confirmation in Bayesian terms. Here these developments can be covered only briefly, and we refer to the entry on confirmation for further details and references.

Statistics has come to play an increasingly important role in the methodology of the experimental sciences from the 19 th century onwards. At that time, statistics and probability theory took on a methodological role as an analysis of inductive inference, and attempts to ground the rationality of induction in the axioms of probability theory have continued throughout the 20 th century and in to the present. Developments in the theory of statistics itself, meanwhile, have had a direct and immense influence on the experimental method, including methods for measuring the uncertainty of observations such as the Method of Least Squares developed by Legendre and Gauss in the early 19 th century, criteria for the rejection of outliers proposed by Peirce by the mid-19 th century, and the significance tests developed by Gosset (a.k.a. “Student”), Fisher, Neyman & Pearson and others in the 1920s and 1930s (see, e.g., Swijtink 1987 for a brief historical overview; and also the entry on C.S. Peirce ).

These developments within statistics then in turn led to a reflective discussion among both statisticians and philosophers of science on how to perceive the process of hypothesis testing: whether it was a rigorous statistical inference that could provide a numerical expression of the degree of confidence in the tested hypothesis, or if it should be seen as a decision between different courses of actions that also involved a value component. This led to a major controversy among Fisher on the one side and Neyman and Pearson on the other (see especially Fisher 1955, Neyman 1956 and Pearson 1955, and for analyses of the controversy, e.g., Howie 2002, Marks 2000, Lenhard 2006). On Fisher’s view, hypothesis testing was a methodology for when to accept or reject a statistical hypothesis, namely that a hypothesis should be rejected by evidence if this evidence would be unlikely relative to other possible outcomes, given the hypothesis were true. In contrast, on Neyman and Pearson’s view, the consequence of error also had to play a role when deciding between hypotheses. Introducing the distinction between the error of rejecting a true hypothesis (type I error) and accepting a false hypothesis (type II error), they argued that it depends on the consequences of the error to decide whether it is more important to avoid rejecting a true hypothesis or accepting a false one. Hence, Fisher aimed for a theory of inductive inference that enabled a numerical expression of confidence in a hypothesis. To him, the important point was the search for truth, not utility. In contrast, the Neyman-Pearson approach provided a strategy of inductive behaviour for deciding between different courses of action. Here, the important point was not whether a hypothesis was true, but whether one should act as if it was.

Similar discussions are found in the philosophical literature. On the one side, Churchman (1948) and Rudner (1953) argued that because scientific hypotheses can never be completely verified, a complete analysis of the methods of scientific inference includes ethical judgments in which the scientists must decide whether the evidence is sufficiently strong or that the probability is sufficiently high to warrant the acceptance of the hypothesis, which again will depend on the importance of making a mistake in accepting or rejecting the hypothesis. Others, such as Jeffrey (1956) and Levi (1960) disagreed and instead defended a value-neutral view of science on which scientists should bracket their attitudes, preferences, temperament, and values when assessing the correctness of their inferences. For more details on this value-free ideal in the philosophy of science and its historical development, see Douglas (2009) and Howard (2003). For a broad set of case studies examining the role of values in science, see e.g. Elliott & Richards 2017.

In recent decades, philosophical discussions of the evaluation of probabilistic hypotheses by statistical inference have largely focused on Bayesianism that understands probability as a measure of a person’s degree of belief in an event, given the available information, and frequentism that instead understands probability as a long-run frequency of a repeatable event. Hence, for Bayesians probabilities refer to a state of knowledge, whereas for frequentists probabilities refer to frequencies of events (see, e.g., Sober 2008, chapter 1 for a detailed introduction to Bayesianism and frequentism as well as to likelihoodism). Bayesianism aims at providing a quantifiable, algorithmic representation of belief revision, where belief revision is a function of prior beliefs (i.e., background knowledge) and incoming evidence. Bayesianism employs a rule based on Bayes’ theorem, a theorem of the probability calculus which relates conditional probabilities. The probability that a particular hypothesis is true is interpreted as a degree of belief, or credence, of the scientist. There will also be a probability and a degree of belief that a hypothesis will be true conditional on a piece of evidence (an observation, say) being true. Bayesianism proscribes that it is rational for the scientist to update their belief in the hypothesis to that conditional probability should it turn out that the evidence is, in fact, observed (see, e.g., Sprenger & Hartmann 2019 for a comprehensive treatment of Bayesian philosophy of science). Originating in the work of Neyman and Person, frequentism aims at providing the tools for reducing long-run error rates, such as the error-statistical approach developed by Mayo (1996) that focuses on how experimenters can avoid both type I and type II errors by building up a repertoire of procedures that detect errors if and only if they are present. Both Bayesianism and frequentism have developed over time, they are interpreted in different ways by its various proponents, and their relations to previous criticism to attempts at defining scientific method are seen differently by proponents and critics. The literature, surveys, reviews and criticism in this area are vast and the reader is referred to the entries on Bayesian epistemology and confirmation .

5. Method in Practice

Attention to scientific practice, as we have seen, is not itself new. However, the turn to practice in the philosophy of science of late can be seen as a correction to the pessimism with respect to method in philosophy of science in later parts of the 20 th century, and as an attempted reconciliation between sociological and rationalist explanations of scientific knowledge. Much of this work sees method as detailed and context specific problem-solving procedures, and methodological analyses to be at the same time descriptive, critical and advisory (see Nickles 1987 for an exposition of this view). The following section contains a survey of some of the practice focuses. In this section we turn fully to topics rather than chronology.

A problem with the distinction between the contexts of discovery and justification that figured so prominently in philosophy of science in the first half of the 20 th century (see section 2 ) is that no such distinction can be clearly seen in scientific activity (see Arabatzis 2006). Thus, in recent decades, it has been recognized that study of conceptual innovation and change should not be confined to psychology and sociology of science, but are also important aspects of scientific practice which philosophy of science should address (see also the entry on scientific discovery ). Looking for the practices that drive conceptual innovation has led philosophers to examine both the reasoning practices of scientists and the wide realm of experimental practices that are not directed narrowly at testing hypotheses, that is, exploratory experimentation.

Examining the reasoning practices of historical and contemporary scientists, Nersessian (2008) has argued that new scientific concepts are constructed as solutions to specific problems by systematic reasoning, and that of analogy, visual representation and thought-experimentation are among the important reasoning practices employed. These ubiquitous forms of reasoning are reliable—but also fallible—methods of conceptual development and change. On her account, model-based reasoning consists of cycles of construction, simulation, evaluation and adaption of models that serve as interim interpretations of the target problem to be solved. Often, this process will lead to modifications or extensions, and a new cycle of simulation and evaluation. However, Nersessian also emphasizes that

creative model-based reasoning cannot be applied as a simple recipe, is not always productive of solutions, and even its most exemplary usages can lead to incorrect solutions. (Nersessian 2008: 11)

Thus, while on the one hand she agrees with many previous philosophers that there is no logic of discovery, discoveries can derive from reasoned processes, such that a large and integral part of scientific practice is

the creation of concepts through which to comprehend, structure, and communicate about physical phenomena …. (Nersessian 1987: 11)

Similarly, work on heuristics for discovery and theory construction by scholars such as Darden (1991) and Bechtel & Richardson (1993) present science as problem solving and investigate scientific problem solving as a special case of problem-solving in general. Drawing largely on cases from the biological sciences, much of their focus has been on reasoning strategies for the generation, evaluation, and revision of mechanistic explanations of complex systems.

Addressing another aspect of the context distinction, namely the traditional view that the primary role of experiments is to test theoretical hypotheses according to the H-D model, other philosophers of science have argued for additional roles that experiments can play. The notion of exploratory experimentation was introduced to describe experiments driven by the desire to obtain empirical regularities and to develop concepts and classifications in which these regularities can be described (Steinle 1997, 2002; Burian 1997; Waters 2007)). However the difference between theory driven experimentation and exploratory experimentation should not be seen as a sharp distinction. Theory driven experiments are not always directed at testing hypothesis, but may also be directed at various kinds of fact-gathering, such as determining numerical parameters. Vice versa , exploratory experiments are usually informed by theory in various ways and are therefore not theory-free. Instead, in exploratory experiments phenomena are investigated without first limiting the possible outcomes of the experiment on the basis of extant theory about the phenomena.

The development of high throughput instrumentation in molecular biology and neighbouring fields has given rise to a special type of exploratory experimentation that collects and analyses very large amounts of data, and these new ‘omics’ disciplines are often said to represent a break with the ideal of hypothesis-driven science (Burian 2007; Elliott 2007; Waters 2007; O’Malley 2007) and instead described as data-driven research (Leonelli 2012; Strasser 2012) or as a special kind of “convenience experimentation” in which many experiments are done simply because they are extraordinarily convenient to perform (Krohs 2012).

5.2 Computer methods and ‘new ways’ of doing science

The field of omics just described is possible because of the ability of computers to process, in a reasonable amount of time, the huge quantities of data required. Computers allow for more elaborate experimentation (higher speed, better filtering, more variables, sophisticated coordination and control), but also, through modelling and simulations, might constitute a form of experimentation themselves. Here, too, we can pose a version of the general question of method versus practice: does the practice of using computers fundamentally change scientific method, or merely provide a more efficient means of implementing standard methods?

Because computers can be used to automate measurements, quantifications, calculations, and statistical analyses where, for practical reasons, these operations cannot be otherwise carried out, many of the steps involved in reaching a conclusion on the basis of an experiment are now made inside a “black box”, without the direct involvement or awareness of a human. This has epistemological implications, regarding what we can know, and how we can know it. To have confidence in the results, computer methods are therefore subjected to tests of verification and validation.

The distinction between verification and validation is easiest to characterize in the case of computer simulations. In a typical computer simulation scenario computers are used to numerically integrate differential equations for which no analytic solution is available. The equations are part of the model the scientist uses to represent a phenomenon or system under investigation. Verifying a computer simulation means checking that the equations of the model are being correctly approximated. Validating a simulation means checking that the equations of the model are adequate for the inferences one wants to make on the basis of that model.

A number of issues related to computer simulations have been raised. The identification of validity and verification as the testing methods has been criticized. Oreskes et al. (1994) raise concerns that “validiation”, because it suggests deductive inference, might lead to over-confidence in the results of simulations. The distinction itself is probably too clean, since actual practice in the testing of simulations mixes and moves back and forth between the two (Weissart 1997; Parker 2008a; Winsberg 2010). Computer simulations do seem to have a non-inductive character, given that the principles by which they operate are built in by the programmers, and any results of the simulation follow from those in-built principles in such a way that those results could, in principle, be deduced from the program code and its inputs. The status of simulations as experiments has therefore been examined (Kaufmann and Smarr 1993; Humphreys 1995; Hughes 1999; Norton and Suppe 2001). This literature considers the epistemology of these experiments: what we can learn by simulation, and also the kinds of justifications which can be given in applying that knowledge to the “real” world. (Mayo 1996; Parker 2008b). As pointed out, part of the advantage of computer simulation derives from the fact that huge numbers of calculations can be carried out without requiring direct observation by the experimenter/​simulator. At the same time, many of these calculations are approximations to the calculations which would be performed first-hand in an ideal situation. Both factors introduce uncertainties into the inferences drawn from what is observed in the simulation.

For many of the reasons described above, computer simulations do not seem to belong clearly to either the experimental or theoretical domain. Rather, they seem to crucially involve aspects of both. This has led some authors, such as Fox Keller (2003: 200) to argue that we ought to consider computer simulation a “qualitatively different way of doing science”. The literature in general tends to follow Kaufmann and Smarr (1993) in referring to computer simulation as a “third way” for scientific methodology (theoretical reasoning and experimental practice are the first two ways.). It should also be noted that the debates around these issues have tended to focus on the form of computer simulation typical in the physical sciences, where models are based on dynamical equations. Other forms of simulation might not have the same problems, or have problems of their own (see the entry on computer simulations in science ).

In recent years, the rapid development of machine learning techniques has prompted some scholars to suggest that the scientific method has become “obsolete” (Anderson 2008, Carrol and Goodstein 2009). This has resulted in an intense debate on the relative merit of data-driven and hypothesis-driven research (for samples, see e.g. Mazzocchi 2015 or Succi and Coveney 2018). For a detailed treatment of this topic, we refer to the entry scientific research and big data .

6. Discourse on scientific method

Despite philosophical disagreements, the idea of the scientific method still figures prominently in contemporary discourse on many different topics, both within science and in society at large. Often, reference to scientific method is used in ways that convey either the legend of a single, universal method characteristic of all science, or grants to a particular method or set of methods privilege as a special ‘gold standard’, often with reference to particular philosophers to vindicate the claims. Discourse on scientific method also typically arises when there is a need to distinguish between science and other activities, or for justifying the special status conveyed to science. In these areas, the philosophical attempts at identifying a set of methods characteristic for scientific endeavors are closely related to the philosophy of science’s classical problem of demarcation (see the entry on science and pseudo-science ) and to the philosophical analysis of the social dimension of scientific knowledge and the role of science in democratic society.

One of the settings in which the legend of a single, universal scientific method has been particularly strong is science education (see, e.g., Bauer 1992; McComas 1996; Wivagg & Allchin 2002). [ 5 ] Often, ‘the scientific method’ is presented in textbooks and educational web pages as a fixed four or five step procedure starting from observations and description of a phenomenon and progressing over formulation of a hypothesis which explains the phenomenon, designing and conducting experiments to test the hypothesis, analyzing the results, and ending with drawing a conclusion. Such references to a universal scientific method can be found in educational material at all levels of science education (Blachowicz 2009), and numerous studies have shown that the idea of a general and universal scientific method often form part of both students’ and teachers’ conception of science (see, e.g., Aikenhead 1987; Osborne et al. 2003). In response, it has been argued that science education need to focus more on teaching about the nature of science, although views have differed on whether this is best done through student-led investigations, contemporary cases, or historical cases (Allchin, Andersen & Nielsen 2014)

Although occasionally phrased with reference to the H-D method, important historical roots of the legend in science education of a single, universal scientific method are the American philosopher and psychologist Dewey’s account of inquiry in How We Think (1910) and the British mathematician Karl Pearson’s account of science in Grammar of Science (1892). On Dewey’s account, inquiry is divided into the five steps of

(i) a felt difficulty, (ii) its location and definition, (iii) suggestion of a possible solution, (iv) development by reasoning of the bearing of the suggestions, (v) further observation and experiment leading to its acceptance or rejection. (Dewey 1910: 72)

Similarly, on Pearson’s account, scientific investigations start with measurement of data and observation of their correction and sequence from which scientific laws can be discovered with the aid of creative imagination. These laws have to be subject to criticism, and their final acceptance will have equal validity for “all normally constituted minds”. Both Dewey’s and Pearson’s accounts should be seen as generalized abstractions of inquiry and not restricted to the realm of science—although both Dewey and Pearson referred to their respective accounts as ‘the scientific method’.

Occasionally, scientists make sweeping statements about a simple and distinct scientific method, as exemplified by Feynman’s simplified version of a conjectures and refutations method presented, for example, in the last of his 1964 Cornell Messenger lectures. [ 6 ] However, just as often scientists have come to the same conclusion as recent philosophy of science that there is not any unique, easily described scientific method. For example, the physicist and Nobel Laureate Weinberg described in the paper “The Methods of Science … And Those By Which We Live” (1995) how

The fact that the standards of scientific success shift with time does not only make the philosophy of science difficult; it also raises problems for the public understanding of science. We do not have a fixed scientific method to rally around and defend. (1995: 8)

Interview studies with scientists on their conception of method shows that scientists often find it hard to figure out whether available evidence confirms their hypothesis, and that there are no direct translations between general ideas about method and specific strategies to guide how research is conducted (Schickore & Hangel 2019, Hangel & Schickore 2017)

Reference to the scientific method has also often been used to argue for the scientific nature or special status of a particular activity. Philosophical positions that argue for a simple and unique scientific method as a criterion of demarcation, such as Popperian falsification, have often attracted practitioners who felt that they had a need to defend their domain of practice. For example, references to conjectures and refutation as the scientific method are abundant in much of the literature on complementary and alternative medicine (CAM)—alongside the competing position that CAM, as an alternative to conventional biomedicine, needs to develop its own methodology different from that of science.

Also within mainstream science, reference to the scientific method is used in arguments regarding the internal hierarchy of disciplines and domains. A frequently seen argument is that research based on the H-D method is superior to research based on induction from observations because in deductive inferences the conclusion follows necessarily from the premises. (See, e.g., Parascandola 1998 for an analysis of how this argument has been made to downgrade epidemiology compared to the laboratory sciences.) Similarly, based on an examination of the practices of major funding institutions such as the National Institutes of Health (NIH), the National Science Foundation (NSF) and the Biomedical Sciences Research Practices (BBSRC) in the UK, O’Malley et al. (2009) have argued that funding agencies seem to have a tendency to adhere to the view that the primary activity of science is to test hypotheses, while descriptive and exploratory research is seen as merely preparatory activities that are valuable only insofar as they fuel hypothesis-driven research.

In some areas of science, scholarly publications are structured in a way that may convey the impression of a neat and linear process of inquiry from stating a question, devising the methods by which to answer it, collecting the data, to drawing a conclusion from the analysis of data. For example, the codified format of publications in most biomedical journals known as the IMRAD format (Introduction, Method, Results, Analysis, Discussion) is explicitly described by the journal editors as “not an arbitrary publication format but rather a direct reflection of the process of scientific discovery” (see the so-called “Vancouver Recommendations”, ICMJE 2013: 11). However, scientific publications do not in general reflect the process by which the reported scientific results were produced. For example, under the provocative title “Is the scientific paper a fraud?”, Medawar argued that scientific papers generally misrepresent how the results have been produced (Medawar 1963/1996). Similar views have been advanced by philosophers, historians and sociologists of science (Gilbert 1976; Holmes 1987; Knorr-Cetina 1981; Schickore 2008; Suppe 1998) who have argued that scientists’ experimental practices are messy and often do not follow any recognizable pattern. Publications of research results, they argue, are retrospective reconstructions of these activities that often do not preserve the temporal order or the logic of these activities, but are instead often constructed in order to screen off potential criticism (see Schickore 2008 for a review of this work).

Philosophical positions on the scientific method have also made it into the court room, especially in the US where judges have drawn on philosophy of science in deciding when to confer special status to scientific expert testimony. A key case is Daubert vs Merrell Dow Pharmaceuticals (92–102, 509 U.S. 579, 1993). In this case, the Supreme Court argued in its 1993 ruling that trial judges must ensure that expert testimony is reliable, and that in doing this the court must look at the expert’s methodology to determine whether the proffered evidence is actually scientific knowledge. Further, referring to works of Popper and Hempel the court stated that

ordinarily, a key question to be answered in determining whether a theory or technique is scientific knowledge … is whether it can be (and has been) tested. (Justice Blackmun, Daubert v. Merrell Dow Pharmaceuticals; see Other Internet Resources for a link to the opinion)

But as argued by Haack (2005a,b, 2010) and by Foster & Hubner (1999), by equating the question of whether a piece of testimony is reliable with the question whether it is scientific as indicated by a special methodology, the court was producing an inconsistent mixture of Popper’s and Hempel’s philosophies, and this has later led to considerable confusion in subsequent case rulings that drew on the Daubert case (see Haack 2010 for a detailed exposition).

The difficulties around identifying the methods of science are also reflected in the difficulties of identifying scientific misconduct in the form of improper application of the method or methods of science. One of the first and most influential attempts at defining misconduct in science was the US definition from 1989 that defined misconduct as

fabrication, falsification, plagiarism, or other practices that seriously deviate from those that are commonly accepted within the scientific community . (Code of Federal Regulations, part 50, subpart A., August 8, 1989, italics added)

However, the “other practices that seriously deviate” clause was heavily criticized because it could be used to suppress creative or novel science. For example, the National Academy of Science stated in their report Responsible Science (1992) that it

wishes to discourage the possibility that a misconduct complaint could be lodged against scientists based solely on their use of novel or unorthodox research methods. (NAS: 27)

This clause was therefore later removed from the definition. For an entry into the key philosophical literature on conduct in science, see Shamoo & Resnick (2009).

The question of the source of the success of science has been at the core of philosophy since the beginning of modern science. If viewed as a matter of epistemology more generally, scientific method is a part of the entire history of philosophy. Over that time, science and whatever methods its practitioners may employ have changed dramatically. Today, many philosophers have taken up the banners of pluralism or of practice to focus on what are, in effect, fine-grained and contextually limited examinations of scientific method. Others hope to shift perspectives in order to provide a renewed general account of what characterizes the activity we call science.

One such perspective has been offered recently by Hoyningen-Huene (2008, 2013), who argues from the history of philosophy of science that after three lengthy phases of characterizing science by its method, we are now in a phase where the belief in the existence of a positive scientific method has eroded and what has been left to characterize science is only its fallibility. First was a phase from Plato and Aristotle up until the 17 th century where the specificity of scientific knowledge was seen in its absolute certainty established by proof from evident axioms; next was a phase up to the mid-19 th century in which the means to establish the certainty of scientific knowledge had been generalized to include inductive procedures as well. In the third phase, which lasted until the last decades of the 20 th century, it was recognized that empirical knowledge was fallible, but it was still granted a special status due to its distinctive mode of production. But now in the fourth phase, according to Hoyningen-Huene, historical and philosophical studies have shown how “scientific methods with the characteristics as posited in the second and third phase do not exist” (2008: 168) and there is no longer any consensus among philosophers and historians of science about the nature of science. For Hoyningen-Huene, this is too negative a stance, and he therefore urges the question about the nature of science anew. His own answer to this question is that “scientific knowledge differs from other kinds of knowledge, especially everyday knowledge, primarily by being more systematic” (Hoyningen-Huene 2013: 14). Systematicity can have several different dimensions: among them are more systematic descriptions, explanations, predictions, defense of knowledge claims, epistemic connectedness, ideal of completeness, knowledge generation, representation of knowledge and critical discourse. Hence, what characterizes science is the greater care in excluding possible alternative explanations, the more detailed elaboration with respect to data on which predictions are based, the greater care in detecting and eliminating sources of error, the more articulate connections to other pieces of knowledge, etc. On this position, what characterizes science is not that the methods employed are unique to science, but that the methods are more carefully employed.

Another, similar approach has been offered by Haack (2003). She sets off, similar to Hoyningen-Huene, from a dissatisfaction with the recent clash between what she calls Old Deferentialism and New Cynicism. The Old Deferentialist position is that science progressed inductively by accumulating true theories confirmed by empirical evidence or deductively by testing conjectures against basic statements; while the New Cynics position is that science has no epistemic authority and no uniquely rational method and is merely just politics. Haack insists that contrary to the views of the New Cynics, there are objective epistemic standards, and there is something epistemologically special about science, even though the Old Deferentialists pictured this in a wrong way. Instead, she offers a new Critical Commonsensist account on which standards of good, strong, supportive evidence and well-conducted, honest, thorough and imaginative inquiry are not exclusive to the sciences, but the standards by which we judge all inquirers. In this sense, science does not differ in kind from other kinds of inquiry, but it may differ in the degree to which it requires broad and detailed background knowledge and a familiarity with a technical vocabulary that only specialists may possess.

  • Aikenhead, G.S., 1987, “High-school graduates’ beliefs about science-technology-society. III. Characteristics and limitations of scientific knowledge”, Science Education , 71(4): 459–487.
  • Allchin, D., H.M. Andersen and K. Nielsen, 2014, “Complementary Approaches to Teaching Nature of Science: Integrating Student Inquiry, Historical Cases, and Contemporary Cases in Classroom Practice”, Science Education , 98: 461–486.
  • Anderson, C., 2008, “The end of theory: The data deluge makes the scientific method obsolete”, Wired magazine , 16(7): 16–07
  • Arabatzis, T., 2006, “On the inextricability of the context of discovery and the context of justification”, in Revisiting Discovery and Justification , J. Schickore and F. Steinle (eds.), Dordrecht: Springer, pp. 215–230.
  • Barnes, J. (ed.), 1984, The Complete Works of Aristotle, Vols I and II , Princeton: Princeton University Press.
  • Barnes, B. and D. Bloor, 1982, “Relativism, Rationalism, and the Sociology of Knowledge”, in Rationality and Relativism , M. Hollis and S. Lukes (eds.), Cambridge: MIT Press, pp. 1–20.
  • Bauer, H.H., 1992, Scientific Literacy and the Myth of the Scientific Method , Urbana: University of Illinois Press.
  • Bechtel, W. and R.C. Richardson, 1993, Discovering complexity , Princeton, NJ: Princeton University Press.
  • Berkeley, G., 1734, The Analyst in De Motu and The Analyst: A Modern Edition with Introductions and Commentary , D. Jesseph (trans. and ed.), Dordrecht: Kluwer Academic Publishers, 1992.
  • Blachowicz, J., 2009, “How science textbooks treat scientific method: A philosopher’s perspective”, The British Journal for the Philosophy of Science , 60(2): 303–344.
  • Bloor, D., 1991, Knowledge and Social Imagery , Chicago: University of Chicago Press, 2 nd edition.
  • Boyle, R., 1682, New experiments physico-mechanical, touching the air , Printed by Miles Flesher for Richard Davis, bookseller in Oxford.
  • Bridgman, P.W., 1927, The Logic of Modern Physics , New York: Macmillan.
  • –––, 1956, “The Methodological Character of Theoretical Concepts”, in The Foundations of Science and the Concepts of Science and Psychology , Herbert Feigl and Michael Scriven (eds.), Minnesota: University of Minneapolis Press, pp. 38–76.
  • Burian, R., 1997, “Exploratory Experimentation and the Role of Histochemical Techniques in the Work of Jean Brachet, 1938–1952”, History and Philosophy of the Life Sciences , 19(1): 27–45.
  • –––, 2007, “On microRNA and the need for exploratory experimentation in post-genomic molecular biology”, History and Philosophy of the Life Sciences , 29(3): 285–311.
  • Carnap, R., 1928, Der logische Aufbau der Welt , Berlin: Bernary, transl. by R.A. George, The Logical Structure of the World , Berkeley: University of California Press, 1967.
  • –––, 1956, “The methodological character of theoretical concepts”, Minnesota studies in the philosophy of science , 1: 38–76.
  • Carrol, S., and D. Goodstein, 2009, “Defining the scientific method”, Nature Methods , 6: 237.
  • Churchman, C.W., 1948, “Science, Pragmatics, Induction”, Philosophy of Science , 15(3): 249–268.
  • Cooper, J. (ed.), 1997, Plato: Complete Works , Indianapolis: Hackett.
  • Darden, L., 1991, Theory Change in Science: Strategies from Mendelian Genetics , Oxford: Oxford University Press
  • Dewey, J., 1910, How we think , New York: Dover Publications (reprinted 1997).
  • Douglas, H., 2009, Science, Policy, and the Value-Free Ideal , Pittsburgh: University of Pittsburgh Press.
  • Dupré, J., 2004, “Miracle of Monism ”, in Naturalism in Question , Mario De Caro and David Macarthur (eds.), Cambridge, MA: Harvard University Press, pp. 36–58.
  • Elliott, K.C., 2007, “Varieties of exploratory experimentation in nanotoxicology”, History and Philosophy of the Life Sciences , 29(3): 311–334.
  • Elliott, K. C., and T. Richards (eds.), 2017, Exploring inductive risk: Case studies of values in science , Oxford: Oxford University Press.
  • Falcon, Andrea, 2005, Aristotle and the science of nature: Unity without uniformity , Cambridge: Cambridge University Press.
  • Feyerabend, P., 1978, Science in a Free Society , London: New Left Books
  • –––, 1988, Against Method , London: Verso, 2 nd edition.
  • Fisher, R.A., 1955, “Statistical Methods and Scientific Induction”, Journal of The Royal Statistical Society. Series B (Methodological) , 17(1): 69–78.
  • Foster, K. and P.W. Huber, 1999, Judging Science. Scientific Knowledge and the Federal Courts , Cambridge: MIT Press.
  • Fox Keller, E., 2003, “Models, Simulation, and ‘computer experiments’”, in The Philosophy of Scientific Experimentation , H. Radder (ed.), Pittsburgh: Pittsburgh University Press, 198–215.
  • Gilbert, G., 1976, “The transformation of research findings into scientific knowledge”, Social Studies of Science , 6: 281–306.
  • Gimbel, S., 2011, Exploring the Scientific Method , Chicago: University of Chicago Press.
  • Goodman, N., 1965, Fact , Fiction, and Forecast , Indianapolis: Bobbs-Merrill.
  • Haack, S., 1995, “Science is neither sacred nor a confidence trick”, Foundations of Science , 1(3): 323–335.
  • –––, 2003, Defending science—within reason , Amherst: Prometheus.
  • –––, 2005a, “Disentangling Daubert: an epistemological study in theory and practice”, Journal of Philosophy, Science and Law , 5, Haack 2005a available online . doi:10.5840/jpsl2005513
  • –––, 2005b, “Trial and error: The Supreme Court’s philosophy of science”, American Journal of Public Health , 95: S66-S73.
  • –––, 2010, “Federal Philosophy of Science: A Deconstruction-and a Reconstruction”, NYUJL & Liberty , 5: 394.
  • Hangel, N. and J. Schickore, 2017, “Scientists’ conceptions of good research practice”, Perspectives on Science , 25(6): 766–791
  • Harper, W.L., 2011, Isaac Newton’s Scientific Method: Turning Data into Evidence about Gravity and Cosmology , Oxford: Oxford University Press.
  • Hempel, C., 1950, “Problems and Changes in the Empiricist Criterion of Meaning”, Revue Internationale de Philosophie , 41(11): 41–63.
  • –––, 1951, “The Concept of Cognitive Significance: A Reconsideration”, Proceedings of the American Academy of Arts and Sciences , 80(1): 61–77.
  • –––, 1965, Aspects of scientific explanation and other essays in the philosophy of science , New York–London: Free Press.
  • –––, 1966, Philosophy of Natural Science , Englewood Cliffs: Prentice-Hall.
  • Holmes, F.L., 1987, “Scientific writing and scientific discovery”, Isis , 78(2): 220–235.
  • Howard, D., 2003, “Two left turns make a right: On the curious political career of North American philosophy of science at midcentury”, in Logical Empiricism in North America , G.L. Hardcastle & A.W. Richardson (eds.), Minneapolis: University of Minnesota Press, pp. 25–93.
  • Hoyningen-Huene, P., 2008, “Systematicity: The nature of science”, Philosophia , 36(2): 167–180.
  • –––, 2013, Systematicity. The Nature of Science , Oxford: Oxford University Press.
  • Howie, D., 2002, Interpreting probability: Controversies and developments in the early twentieth century , Cambridge: Cambridge University Press.
  • Hughes, R., 1999, “The Ising Model, Computer Simulation, and Universal Physics”, in Models as Mediators , M. Morgan and M. Morrison (eds.), Cambridge: Cambridge University Press, pp. 97–145
  • Hume, D., 1739, A Treatise of Human Nature , D. Fate Norton and M.J. Norton (eds.), Oxford: Oxford University Press, 2000.
  • Humphreys, P., 1995, “Computational science and scientific method”, Minds and Machines , 5(1): 499–512.
  • ICMJE, 2013, “Recommendations for the Conduct, Reporting, Editing, and Publication of Scholarly Work in Medical Journals”, International Committee of Medical Journal Editors, available online , accessed August 13 2014
  • Jeffrey, R.C., 1956, “Valuation and Acceptance of Scientific Hypotheses”, Philosophy of Science , 23(3): 237–246.
  • Kaufmann, W.J., and L.L. Smarr, 1993, Supercomputing and the Transformation of Science , New York: Scientific American Library.
  • Knorr-Cetina, K., 1981, The Manufacture of Knowledge , Oxford: Pergamon Press.
  • Krohs, U., 2012, “Convenience experimentation”, Studies in History and Philosophy of Biological and BiomedicalSciences , 43: 52–57.
  • Kuhn, T.S., 1962, The Structure of Scientific Revolutions , Chicago: University of Chicago Press
  • Latour, B. and S. Woolgar, 1986, Laboratory Life: The Construction of Scientific Facts , Princeton: Princeton University Press, 2 nd edition.
  • Laudan, L., 1968, “Theories of scientific method from Plato to Mach”, History of Science , 7(1): 1–63.
  • Lenhard, J., 2006, “Models and statistical inference: The controversy between Fisher and Neyman-Pearson”, The British Journal for the Philosophy of Science , 57(1): 69–91.
  • Leonelli, S., 2012, “Making Sense of Data-Driven Research in the Biological and the Biomedical Sciences”, Studies in the History and Philosophy of the Biological and Biomedical Sciences , 43(1): 1–3.
  • Levi, I., 1960, “Must the scientist make value judgments?”, Philosophy of Science , 57(11): 345–357
  • Lindley, D., 1991, Theory Change in Science: Strategies from Mendelian Genetics , Oxford: Oxford University Press.
  • Lipton, P., 2004, Inference to the Best Explanation , London: Routledge, 2 nd edition.
  • Marks, H.M., 2000, The progress of experiment: science and therapeutic reform in the United States, 1900–1990 , Cambridge: Cambridge University Press.
  • Mazzochi, F., 2015, “Could Big Data be the end of theory in science?”, EMBO reports , 16: 1250–1255.
  • Mayo, D.G., 1996, Error and the Growth of Experimental Knowledge , Chicago: University of Chicago Press.
  • McComas, W.F., 1996, “Ten myths of science: Reexamining what we think we know about the nature of science”, School Science and Mathematics , 96(1): 10–16.
  • Medawar, P.B., 1963/1996, “Is the scientific paper a fraud”, in The Strange Case of the Spotted Mouse and Other Classic Essays on Science , Oxford: Oxford University Press, 33–39.
  • Mill, J.S., 1963, Collected Works of John Stuart Mill , J. M. Robson (ed.), Toronto: University of Toronto Press
  • NAS, 1992, Responsible Science: Ensuring the integrity of the research process , Washington DC: National Academy Press.
  • Nersessian, N.J., 1987, “A cognitive-historical approach to meaning in scientific theories”, in The process of science , N. Nersessian (ed.), Berlin: Springer, pp. 161–177.
  • –––, 2008, Creating Scientific Concepts , Cambridge: MIT Press.
  • Newton, I., 1726, Philosophiae naturalis Principia Mathematica (3 rd edition), in The Principia: Mathematical Principles of Natural Philosophy: A New Translation , I.B. Cohen and A. Whitman (trans.), Berkeley: University of California Press, 1999.
  • –––, 1704, Opticks or A Treatise of the Reflections, Refractions, Inflections & Colors of Light , New York: Dover Publications, 1952.
  • Neyman, J., 1956, “Note on an Article by Sir Ronald Fisher”, Journal of the Royal Statistical Society. Series B (Methodological) , 18: 288–294.
  • Nickles, T., 1987, “Methodology, heuristics, and rationality”, in Rational changes in science: Essays on Scientific Reasoning , J.C. Pitt (ed.), Berlin: Springer, pp. 103–132.
  • Nicod, J., 1924, Le problème logique de l’induction , Paris: Alcan. (Engl. transl. “The Logical Problem of Induction”, in Foundations of Geometry and Induction , London: Routledge, 2000.)
  • Nola, R. and H. Sankey, 2000a, “A selective survey of theories of scientific method”, in Nola and Sankey 2000b: 1–65.
  • –––, 2000b, After Popper, Kuhn and Feyerabend. Recent Issues in Theories of Scientific Method , London: Springer.
  • –––, 2007, Theories of Scientific Method , Stocksfield: Acumen.
  • Norton, S., and F. Suppe, 2001, “Why atmospheric modeling is good science”, in Changing the Atmosphere: Expert Knowledge and Environmental Governance , C. Miller and P. Edwards (eds.), Cambridge, MA: MIT Press, 88–133.
  • O’Malley, M., 2007, “Exploratory experimentation and scientific practice: Metagenomics and the proteorhodopsin case”, History and Philosophy of the Life Sciences , 29(3): 337–360.
  • O’Malley, M., C. Haufe, K. Elliot, and R. Burian, 2009, “Philosophies of Funding”, Cell , 138: 611–615.
  • Oreskes, N., K. Shrader-Frechette, and K. Belitz, 1994, “Verification, Validation and Confirmation of Numerical Models in the Earth Sciences”, Science , 263(5147): 641–646.
  • Osborne, J., S. Simon, and S. Collins, 2003, “Attitudes towards science: a review of the literature and its implications”, International Journal of Science Education , 25(9): 1049–1079.
  • Parascandola, M., 1998, “Epidemiology—2 nd -Rate Science”, Public Health Reports , 113(4): 312–320.
  • Parker, W., 2008a, “Franklin, Holmes and the Epistemology of Computer Simulation”, International Studies in the Philosophy of Science , 22(2): 165–83.
  • –––, 2008b, “Computer Simulation through an Error-Statistical Lens”, Synthese , 163(3): 371–84.
  • Pearson, K. 1892, The Grammar of Science , London: J.M. Dents and Sons, 1951
  • Pearson, E.S., 1955, “Statistical Concepts in Their Relation to Reality”, Journal of the Royal Statistical Society , B, 17: 204–207.
  • Pickering, A., 1984, Constructing Quarks: A Sociological History of Particle Physics , Edinburgh: Edinburgh University Press.
  • Popper, K.R., 1959, The Logic of Scientific Discovery , London: Routledge, 2002
  • –––, 1963, Conjectures and Refutations , London: Routledge, 2002.
  • –––, 1985, Unended Quest: An Intellectual Autobiography , La Salle: Open Court Publishing Co..
  • Rudner, R., 1953, “The Scientist Qua Scientist Making Value Judgments”, Philosophy of Science , 20(1): 1–6.
  • Rudolph, J.L., 2005, “Epistemology for the masses: The origin of ‘The Scientific Method’ in American Schools”, History of Education Quarterly , 45(3): 341–376
  • Schickore, J., 2008, “Doing science, writing science”, Philosophy of Science , 75: 323–343.
  • Schickore, J. and N. Hangel, 2019, “‘It might be this, it should be that…’ uncertainty and doubt in day-to-day science practice”, European Journal for Philosophy of Science , 9(2): 31. doi:10.1007/s13194-019-0253-9
  • Shamoo, A.E. and D.B. Resnik, 2009, Responsible Conduct of Research , Oxford: Oxford University Press.
  • Shank, J.B., 2008, The Newton Wars and the Beginning of the French Enlightenment , Chicago: The University of Chicago Press.
  • Shapin, S. and S. Schaffer, 1985, Leviathan and the air-pump , Princeton: Princeton University Press.
  • Smith, G.E., 2002, “The Methodology of the Principia”, in The Cambridge Companion to Newton , I.B. Cohen and G.E. Smith (eds.), Cambridge: Cambridge University Press, 138–173.
  • Snyder, L.J., 1997a, “Discoverers’ Induction”, Philosophy of Science , 64: 580–604.
  • –––, 1997b, “The Mill-Whewell Debate: Much Ado About Induction”, Perspectives on Science , 5: 159–198.
  • –––, 1999, “Renovating the Novum Organum: Bacon, Whewell and Induction”, Studies in History and Philosophy of Science , 30: 531–557.
  • Sober, E., 2008, Evidence and Evolution. The logic behind the science , Cambridge: Cambridge University Press
  • Sprenger, J. and S. Hartmann, 2019, Bayesian philosophy of science , Oxford: Oxford University Press.
  • Steinle, F., 1997, “Entering New Fields: Exploratory Uses of Experimentation”, Philosophy of Science (Proceedings), 64: S65–S74.
  • –––, 2002, “Experiments in History and Philosophy of Science”, Perspectives on Science , 10(4): 408–432.
  • Strasser, B.J., 2012, “Data-driven sciences: From wonder cabinets to electronic databases”, Studies in History and Philosophy of Science Part C: Studies in History and Philosophy of Biological and Biomedical Sciences , 43(1): 85–87.
  • Succi, S. and P.V. Coveney, 2018, “Big data: the end of the scientific method?”, Philosophical Transactions of the Royal Society A , 377: 20180145. doi:10.1098/rsta.2018.0145
  • Suppe, F., 1998, “The Structure of a Scientific Paper”, Philosophy of Science , 65(3): 381–405.
  • Swijtink, Z.G., 1987, “The objectification of observation: Measurement and statistical methods in the nineteenth century”, in The probabilistic revolution. Ideas in History, Vol. 1 , L. Kruger (ed.), Cambridge MA: MIT Press, pp. 261–285.
  • Waters, C.K., 2007, “The nature and context of exploratory experimentation: An introduction to three case studies of exploratory research”, History and Philosophy of the Life Sciences , 29(3): 275–284.
  • Weinberg, S., 1995, “The methods of science… and those by which we live”, Academic Questions , 8(2): 7–13.
  • Weissert, T., 1997, The Genesis of Simulation in Dynamics: Pursuing the Fermi-Pasta-Ulam Problem , New York: Springer Verlag.
  • William H., 1628, Exercitatio Anatomica de Motu Cordis et Sanguinis in Animalibus , in On the Motion of the Heart and Blood in Animals , R. Willis (trans.), Buffalo: Prometheus Books, 1993.
  • Winsberg, E., 2010, Science in the Age of Computer Simulation , Chicago: University of Chicago Press.
  • Wivagg, D. & D. Allchin, 2002, “The Dogma of the Scientific Method”, The American Biology Teacher , 64(9): 645–646
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.
  • Blackmun opinion , in Daubert v. Merrell Dow Pharmaceuticals (92–102), 509 U.S. 579 (1993).
  • Scientific Method at philpapers. Darrell Rowbottom (ed.).
  • Recent Articles | Scientific Method | The Scientist Magazine

al-Kindi | Albert the Great [= Albertus magnus] | Aquinas, Thomas | Arabic and Islamic Philosophy, disciplines in: natural philosophy and natural science | Arabic and Islamic Philosophy, historical and methodological topics in: Greek sources | Arabic and Islamic Philosophy, historical and methodological topics in: influence of Arabic and Islamic Philosophy on the Latin West | Aristotle | Bacon, Francis | Bacon, Roger | Berkeley, George | biology: experiment in | Boyle, Robert | Cambridge Platonists | confirmation | Descartes, René | Enlightenment | epistemology | epistemology: Bayesian | epistemology: social | Feyerabend, Paul | Galileo Galilei | Grosseteste, Robert | Hempel, Carl | Hume, David | Hume, David: Newtonianism and Anti-Newtonianism | induction: problem of | Kant, Immanuel | Kuhn, Thomas | Leibniz, Gottfried Wilhelm | Locke, John | Mill, John Stuart | More, Henry | Neurath, Otto | Newton, Isaac | Newton, Isaac: philosophy | Ockham [Occam], William | operationalism | Peirce, Charles Sanders | Plato | Popper, Karl | rationality: historicist theories of | Reichenbach, Hans | reproducibility, scientific | Schlick, Moritz | science: and pseudo-science | science: theory and observation in | science: unity of | scientific discovery | scientific knowledge: social dimensions of | simulations in science | skepticism: medieval | space and time: absolute and relational space and motion, post-Newtonian theories | Vienna Circle | Whewell, William | Zabarella, Giacomo

Copyright © 2021 by Brian Hepburn < brian . hepburn @ wichita . edu > Hanne Andersen < hanne . andersen @ ind . ku . dk >

  • 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

loading

Encyclopedia Britannica

  • History & Society
  • Science & Tech
  • Biographies
  • Animals & Nature
  • Geography & Travel
  • Arts & Culture
  • Games & Quizzes
  • On This Day
  • One Good Fact
  • New Articles
  • Lifestyles & Social Issues
  • Philosophy & Religion
  • Politics, Law & Government
  • World History
  • Health & Medicine
  • Browse Biographies
  • Birds, Reptiles & Other Vertebrates
  • Bugs, Mollusks & Other Invertebrates
  • Environment
  • Fossils & Geologic Time
  • Entertainment & Pop Culture
  • Sports & Recreation
  • Visual Arts
  • Demystified
  • Image Galleries
  • Infographics
  • Top Questions
  • Britannica Kids
  • Saving Earth
  • Space Next 50
  • Student Center

flow chart of scientific method

scientific method

Our editors will review what you’ve submitted and determine whether to revise the article.

  • University of Nevada, Reno - College of Agriculture, Biotechnology and Natural Resources Extension - The Scientific Method
  • World History Encyclopedia - Scientific Method
  • LiveScience - What Is Science?
  • Verywell Mind - Scientific Method Steps in Psychology Research
  • WebMD - What is the Scientific Method?
  • Chemistry LibreTexts - The Scientific Method
  • National Center for Biotechnology Information - PubMed Central - Redefining the scientific method: as the use of sophisticated scientific methods that extend our mind
  • Khan Academy - The scientific method
  • Simply Psychology - What are the steps in the Scientific Method?
  • Stanford Encyclopedia of Philosophy - Scientific Method

flow chart of scientific method

scientific method , mathematical and experimental technique employed in the sciences . More specifically, it is the technique used in the construction and testing of a scientific hypothesis .

The process of observing, asking questions, and seeking answers through tests and experiments is not unique to any one field of science. In fact, the scientific method is applied broadly in science, across many different fields. Many empirical sciences, especially the social sciences , use mathematical tools borrowed from probability theory and statistics , together with outgrowths of these, such as decision theory , game theory , utility theory, and operations research . Philosophers of science have addressed general methodological problems, such as the nature of scientific explanation and the justification of induction .

characteristics of scientific method of problem solving

The scientific method is critical to the development of scientific theories , which explain empirical (experiential) laws in a scientifically rational manner. In a typical application of the scientific method, a researcher develops a hypothesis , tests it through various means, and then modifies the hypothesis on the basis of the outcome of the tests and experiments. The modified hypothesis is then retested, further modified, and tested again, until it becomes consistent with observed phenomena and testing outcomes. In this way, hypotheses serve as tools by which scientists gather data. From that data and the many different scientific investigations undertaken to explore hypotheses, scientists are able to develop broad general explanations, or scientific theories.

See also Mill’s methods ; hypothetico-deductive method .

What Are The Steps Of The Scientific Method?

Julia Simkus

Editor at Simply Psychology

BA (Hons) Psychology, Princeton University

Julia Simkus is a graduate of Princeton University with a Bachelor of Arts in Psychology. She is currently studying for a Master's Degree in Counseling for Mental Health and Wellness in September 2023. Julia's research has been published in peer reviewed journals.

Learn about our Editorial Process

Saul McLeod, PhD

Editor-in-Chief for Simply Psychology

BSc (Hons) Psychology, MRes, PhD, University of Manchester

Saul McLeod, PhD., is a qualified psychology teacher with over 18 years of experience in further and higher education. He has been published in peer-reviewed journals, including the Journal of Clinical Psychology.

Olivia Guy-Evans, MSc

Associate Editor for Simply Psychology

BSc (Hons) Psychology, MSc Psychology of Education

Olivia Guy-Evans is a writer and associate editor for Simply Psychology. She has previously worked in healthcare and educational sectors.

On This Page:

Science is not just knowledge. It is also a method for obtaining knowledge. Scientific understanding is organized into theories.

The scientific method is a step-by-step process used by researchers and scientists to determine if there is a relationship between two or more variables. Psychologists use this method to conduct psychological research, gather data, process information, and describe behaviors.

It involves careful observation, asking questions, formulating hypotheses, experimental testing, and refining hypotheses based on experimental findings.

How it is Used

The scientific method can be applied broadly in science across many different fields, such as chemistry, physics, geology, and psychology. In a typical application of this process, a researcher will develop a hypothesis, test this hypothesis, and then modify the hypothesis based on the outcomes of the experiment.

The process is then repeated with the modified hypothesis until the results align with the observed phenomena. Detailed steps of the scientific method are described below.

Keep in mind that the scientific method does not have to follow this fixed sequence of steps; rather, these steps represent a set of general principles or guidelines.

7 Steps of the Scientific Method

Psychology uses an empirical approach.

Empiricism (founded by John Locke) states that the only source of knowledge comes through our senses – e.g., sight, hearing, touch, etc.

Empirical evidence does not rely on argument or belief. Thus, empiricism is the view that all knowledge is based on or may come from direct observation and experience.

The empiricist approach of gaining knowledge through experience quickly became the scientific approach and greatly influenced the development of physics and chemistry in the 17th and 18th centuries.

Steps of the Scientific Method

Step 1: Make an Observation (Theory Construction)

Every researcher starts at the very beginning. Before diving in and exploring something, one must first determine what they will study – it seems simple enough!

By making observations, researchers can establish an area of interest. Once this topic of study has been chosen, a researcher should review existing literature to gain insight into what has already been tested and determine what questions remain unanswered.

This assessment will provide helpful information about what has already been comprehended about the specific topic and what questions remain, and if one can go and answer them.

Specifically, a literature review might implicate examining a substantial amount of documented material from academic journals to books dating back decades. The most appropriate information gathered by the researcher will be shown in the introduction section or abstract of the published study results.

The background material and knowledge will help the researcher with the first significant step in conducting a psychology study, which is formulating a research question.

This is the inductive phase of the scientific process. Observations yield information that is used to formulate theories as explanations. A theory is a well-developed set of ideas that propose an explanation for observed phenomena.

Inductive reasoning moves from specific premises to a general conclusion. It starts with observations of phenomena in the natural world and derives a general law.

Step 2: Ask a Question

Once a researcher has made observations and conducted background research, the next step is to ask a scientific question. A scientific question must be defined, testable, and measurable.

A useful approach to develop a scientific question is: “What is the effect of…?” or “How does X affect Y?”

To answer an experimental question, a researcher must identify two variables: the independent and dependent variables.

The independent variable is the variable manipulated (the cause), and the dependent variable is the variable being measured (the effect).

An example of a research question could be, “Is handwriting or typing more effective for retaining information?” Answering the research question and proposing a relationship between the two variables is discussed in the next step.

Step 3: Form a Hypothesis (Make Predictions)

A hypothesis is an educated guess about the relationship between two or more variables. A hypothesis is an attempt to answer your research question based on prior observation and background research. Theories tend to be too complex to be tested all at once; instead, researchers create hypotheses to test specific aspects of a theory.

For example, a researcher might ask about the connection between sleep and educational performance. Do students who get less sleep perform worse on tests at school?

It is crucial to think about different questions one might have about a particular topic to formulate a reasonable hypothesis. It would help if one also considered how one could investigate the causalities.

It is important that the hypothesis is both testable against reality and falsifiable. This means that it can be tested through an experiment and can be proven wrong.

The falsification principle, proposed by Karl Popper , is a way of demarcating science from non-science. It suggests that for a theory to be considered scientific, it must be able to be tested and conceivably proven false.

To test a hypothesis, we first assume that there is no difference between the populations from which the samples were taken. This is known as the null hypothesis and predicts that the independent variable will not influence the dependent variable.

Examples of “if…then…” Hypotheses:

  • If one gets less than 6 hours of sleep, then one will do worse on tests than if one obtains more rest.
  • If one drinks lots of water before going to bed, one will have to use the bathroom often at night.
  • If one practices exercising and lighting weights, then one’s body will begin to build muscle.

The research hypothesis is often called the alternative hypothesis and predicts what change(s) will occur in the dependent variable when the independent variable is manipulated.

It states that the results are not due to chance and that they are significant in terms of supporting the theory being investigated.

Although one could state and write a scientific hypothesis in many ways, hypotheses are usually built like “if…then…” statements.

Step 4: Run an Experiment (Gather Data)

The next step in the scientific method is to test your hypothesis and collect data. A researcher will design an experiment to test the hypothesis and gather data that will either support or refute the hypothesis.

The exact research methods used to examine a hypothesis depend on what is being studied. A psychologist might utilize two primary forms of research, experimental research, and descriptive research.

The scientific method is objective in that researchers do not let preconceived ideas or biases influence the collection of data and is systematic in that experiments are conducted in a logical way.

Experimental Research

Experimental research is used to investigate cause-and-effect associations between two or more variables. This type of research systematically controls an independent variable and measures its effect on a specified dependent variable.

Experimental research involves manipulating an independent variable and measuring the effect(s) on the dependent variable. Repeating the experiment multiple times is important to confirm that your results are accurate and consistent.

One of the significant advantages of this method is that it permits researchers to determine if changes in one variable cause shifts in each other.

While experiments in psychology typically have many moving parts (and can be relatively complex), an easy investigation is rather fundamental. Still, it does allow researchers to specify cause-and-effect associations between variables.

Most simple experiments use a control group, which involves those who do not receive the treatment, and an experimental group, which involves those who do receive the treatment.

An example of experimental research would be when a pharmaceutical company wants to test a new drug. They give one group a placebo (control group) and the other the actual pill (experimental group).

Descriptive Research

Descriptive research is generally used when it is challenging or even impossible to control the variables in question. Examples of descriptive analysis include naturalistic observation, case studies , and correlation studies .

One example of descriptive research includes phone surveys that marketers often use. While they typically do not allow researchers to identify cause and effect, correlational studies are quite common in psychology research. They make it possible to spot associations between distinct variables and measure the solidity of those relationships.

Step 5: Analyze the Data and Draw Conclusions

Once a researcher has designed and done the investigation and collected sufficient data, it is time to inspect this gathered information and judge what has been found. Researchers can summarize the data, interpret the results, and draw conclusions based on this evidence using analyses and statistics.

Upon completion of the experiment, you can collect your measurements and analyze the data using statistics. Based on the outcomes, you will either reject or confirm your hypothesis.

Analyze the Data

So, how does a researcher determine what the results of their study mean? Statistical analysis can either support or refute a researcher’s hypothesis and can also be used to determine if the conclusions are statistically significant.

When outcomes are said to be “statistically significant,” it is improbable that these results are due to luck or chance. Based on these observations, investigators must then determine what the results mean.

An experiment will support a hypothesis in some circumstances, but sometimes it fails to be truthful in other cases.

What occurs if the developments of a psychology investigation do not endorse the researcher’s hypothesis? It does mean that the study was worthless. Simply because the findings fail to defend the researcher’s hypothesis does not mean that the examination is not helpful or instructive.

This kind of research plays a vital role in supporting scientists in developing unexplored questions and hypotheses to investigate in the future. After decisions have been made, the next step is to communicate the results with the rest of the scientific community.

This is an integral part of the process because it contributes to the general knowledge base and can assist other scientists in finding new research routes to explore.

If the hypothesis is not supported, a researcher should acknowledge the experiment’s results, formulate a new hypothesis, and develop a new experiment.

We must avoid any reference to results proving a theory as this implies 100% certainty, and there is always a chance that evidence may exist that could refute a theory.

Draw Conclusions and Interpret the Data

When the empirical observations disagree with the hypothesis, a number of possibilities must be considered. It might be that the theory is incorrect, in which case it needs altering, so it fully explains the data.

Alternatively, it might be that the hypothesis was poorly derived from the original theory, in which case the scientists were expecting the wrong thing to happen.

It might also be that the research was poorly conducted, or used an inappropriate method, or there were factors in play that the researchers did not consider. This will begin the process of the scientific method again.

If the hypothesis is supported, the researcher can find more evidence to support their hypothesis or look for counter-evidence to strengthen their hypothesis further.

In either scenario, the researcher should share their results with the greater scientific community.

Step 6: Share Your Results

One of the final stages of the research cycle involves the publication of the research. Once the report is written, the researcher(s) may submit the work for publication in an appropriate journal.

Usually, this is done by writing up a study description and publishing the article in a professional or academic journal. The studies and conclusions of psychological work can be seen in peer-reviewed journals such as  Developmental Psychology , Psychological Bulletin, the  Journal of Social Psychology, and numerous others.

Scientists should report their findings by writing up a description of their study and any subsequent findings. This enables other researchers to build upon the present research or replicate the results.

As outlined by the American Psychological Association (APA), there is a typical structure of a journal article that follows a specified format. In these articles, researchers:

  • Supply a brief narrative and background on previous research
  • Give their hypothesis
  • Specify who participated in the study and how they were chosen
  • Provide operational definitions for each variable
  • Explain the measures and methods used to collect data
  • Describe how the data collected was interpreted
  • Discuss what the outcomes mean

A detailed record of psychological studies and all scientific studies is vital to clearly explain the steps and procedures used throughout the study. So that other researchers can try this experiment too and replicate the results.

The editorial process utilized by academic and professional journals guarantees that each submitted article undergoes a thorough peer review to help assure that the study is scientifically sound. Once published, the investigation becomes another piece of the current puzzle of our knowledge “base” on that subject.

This last step is important because all results, whether they supported or did not support the hypothesis, can contribute to the scientific community. Publication of empirical observations leads to more ideas that are tested against the real world, and so on. In this sense, the scientific process is circular.

The editorial process utilized by academic and professional journals guarantees that each submitted article undergoes a thorough peer review to help assure that the study is scientifically sound.

Once published, the investigation becomes another piece of the current puzzle of our knowledge “base” on that subject.

By replicating studies, psychologists can reduce errors, validate theories, and gain a stronger understanding of a particular topic.

Step 7: Repeat the Scientific Method (Iteration)

Now, if one’s hypothesis turns out to be accurate, find more evidence or find counter-evidence. If one’s hypothesis is false, create a new hypothesis or try again.

One may wish to revise their first hypothesis to make a more niche experiment to design or a different specific question to test.

The amazingness of the scientific method is that it is a comprehensive and straightforward process that scientists, and everyone, can utilize over and over again.

So, draw conclusions and repeat because the scientific method is never-ending, and no result is ever considered perfect.

The scientific method is a process of:

  • Making an observation.
  • Forming a hypothesis.
  • Making a prediction.
  • Experimenting to test the hypothesis.

The procedure of repeating the scientific method is crucial to science and all fields of human knowledge.

Further Information

  • Karl Popper – Falsification
  • Thomas – Kuhn Paradigm Shift
  • Positivism in Sociology: Definition, Theory & Examples
  • Is Psychology a Science?
  • Psychology as a Science (PDF)

List the 6 steps of the scientific methods in order

  • Make an observation (theory construction)
  • Ask a question. A scientific question must be defined, testable, and measurable.
  • Form a hypothesis (make predictions)
  • Run an experiment to test the hypothesis (gather data)
  • Analyze the data and draw conclusions
  • Share your results so that other researchers can make new hypotheses

What is the first step of the scientific method?

The first step of the scientific method is making an observation. This involves noticing and describing a phenomenon or group of phenomena that one finds interesting and wishes to explain.

Observations can occur in a natural setting or within the confines of a laboratory. The key point is that the observation provides the initial question or problem that the rest of the scientific method seeks to answer or solve.

What is the scientific method?

The scientific method is a step-by-step process that investigators can follow to determine if there is a causal connection between two or more variables.

Psychologists and other scientists regularly suggest motivations for human behavior. On a more casual level, people judge other people’s intentions, incentives, and actions daily.

While our standard assessments of human behavior are subjective and anecdotal, researchers use the scientific method to study psychology objectively and systematically.

All utilize a scientific method to study distinct aspects of people’s thinking and behavior. This process allows scientists to analyze and understand various psychological phenomena, but it also provides investigators and others a way to disseminate and debate the results of their studies.

The outcomes of these studies are often noted in popular media, which leads numerous to think about how or why researchers came to the findings they did.

Why Use the Six Steps of the Scientific Method

The goal of scientists is to understand better the world that surrounds us. Scientific research is the most critical tool for navigating and learning about our complex world.

Without it, we would be compelled to rely solely on intuition, other people’s power, and luck. We can eliminate our preconceived concepts and superstitions through methodical scientific research and gain an objective sense of ourselves and our world.

All psychological studies aim to explain, predict, and even control or impact mental behaviors or processes. So, psychologists use and repeat the scientific method (and its six steps) to perform and record essential psychological research.

So, psychologists focus on understanding behavior and the cognitive (mental) and physiological (body) processes underlying behavior.

In the real world, people use to understand the behavior of others, such as intuition and personal experience. The hallmark of scientific research is evidence to support a claim.

Scientific knowledge is empirical, meaning it is grounded in objective, tangible evidence that can be observed repeatedly, regardless of who is watching.

The scientific method is crucial because it minimizes the impact of bias or prejudice on the experimenter. Regardless of how hard one tries, even the best-intentioned scientists can’t escape discrimination. can’t

It stems from personal opinions and cultural beliefs, meaning any mortal filters data based on one’s experience. Sadly, this “filtering” process can cause a scientist to favor one outcome over another.

For an everyday person trying to solve a minor issue at home or work, succumbing to these biases is not such a big deal; in fact, most times, it is important.

But in the scientific community, where results must be inspected and reproduced, bias or discrimination must be avoided.

When to Use the Six Steps of the Scientific Method ?

One can use the scientific method anytime, anywhere! From the smallest conundrum to solving global problems, it is a process that can be applied to any science and any investigation.

Even if you are not considered a “scientist,” you will be surprised to know that people of all disciplines use it for all kinds of dilemmas.

Try to catch yourself next time you come by a question and see how you subconsciously or consciously use the scientific method.

Print Friendly, PDF & Email

Scientific Method

Illustration by J.R. Bee. ThoughtCo. 

  • Cell Biology
  • Weather & Climate
  • B.A., Biology, Emory University
  • A.S., Nursing, Chattahoochee Technical College

The scientific method is a series of steps followed by scientific investigators to answer specific questions about the natural world. It involves making observations, formulating a hypothesis , and conducting scientific experiments . Scientific inquiry starts with an observation followed by the formulation of a question about what has been observed. The steps of the scientific method are as follows:

Observation

The first step of the scientific method involves making an observation about something that interests you. This is very important if you are doing a science project because you want your project to be focused on something that will hold your attention. Your observation can be on anything from plant movement to animal behavior, as long as it is something you really want to know more about.​ This is where you come up with the idea for your science project.

Once you've made your observation, you must formulate a question about what you have observed. Your question should tell what it is that you are trying to discover or accomplish in your experiment. When stating your question you should be as specific as possible.​ For example, if you are doing a project on plants , you may want to know how plants interact with microbes. Your question may be: Do plant spices inhibit bacterial growth ?

The hypothesis is a key component of the scientific process. A hypothesis is an idea that is suggested as an explanation for a natural event, a particular experience, or a specific condition that can be tested through definable experimentation. It states the purpose of your experiment, the variables used, and the predicted outcome of your experiment. It is important to note that a hypothesis must be testable. That means that you should be able to test your hypothesis through experimentation .​ Your hypothesis must either be supported or falsified by your experiment. An example of a good hypothesis is: If there is a relation between listening to music and heart rate, then listening to music will cause a person's resting heart rate to either increase or decrease.

Once you've developed a hypothesis, you must design and conduct an experiment that will test it. You should develop a procedure that states very clearly how you plan to conduct your experiment. It is important that you include and identify a controlled variable or dependent variable in your procedure. Controls allow us to test a single variable in an experiment because they are unchanged. We can then make observations and comparisons between our controls and our independent variables (things that change in the experiment) to develop an accurate conclusion.​

The results are where you report what happened in the experiment. That includes detailing all observations and data made during your experiment. Most people find it easier to visualize the data by charting or graphing the information.​

The final step of the scientific method is developing a conclusion. This is where all of the results from the experiment are analyzed and a determination is reached about the hypothesis. Did the experiment support or reject your hypothesis? If your hypothesis was supported, great. If not, repeat the experiment or think of ways to improve your procedure.

  • Biology Prefixes and Suffixes: chrom- or chromo-
  • Biology Prefixes and Suffixes: proto-
  • 6 Things You Should Know About Biological Evolution
  • Biology Prefixes and Suffixes: Aer- or Aero-
  • Taxonomy and Organism Classification
  • Homeostasis
  • Biology Prefixes and Suffixes: diplo-
  • The Biology Suffix -lysis
  • Biology Prefixes and Suffixes Index
  • Biology Prefixes and Suffixes: tel- or telo-
  • Parasitism: Definition and Examples
  • Biology Prefixes and Suffixes: Erythr- or Erythro-
  • Biology Prefixes and Suffixes: ana-
  • Biology Prefixes and Suffixes: phago- or phag-
  • Biology Prefixes and Suffixes: -phyll or -phyl
  • Biology Prefixes and Suffixes: haplo-

PrepScholar

Choose Your Test

  • Search Blogs By Category
  • College Admissions
  • AP and IB Exams
  • GPA and Coursework

The 6 Scientific Method Steps and How to Use Them

author image

General Education

feature_microscope-1

When you’re faced with a scientific problem, solving it can seem like an impossible prospect. There are so many possible explanations for everything we see and experience—how can you possibly make sense of them all? Science has a simple answer: the scientific method.

The scientific method is a method of asking and answering questions about the world. These guiding principles give scientists a model to work through when trying to understand the world, but where did that model come from, and how does it work?

In this article, we’ll define the scientific method, discuss its long history, and cover each of the scientific method steps in detail.

What Is the Scientific Method?

At its most basic, the scientific method is a procedure for conducting scientific experiments. It’s a set model that scientists in a variety of fields can follow, going from initial observation to conclusion in a loose but concrete format.

The number of steps varies, but the process begins with an observation, progresses through an experiment, and concludes with analysis and sharing data. One of the most important pieces to the scientific method is skepticism —the goal is to find truth, not to confirm a particular thought. That requires reevaluation and repeated experimentation, as well as examining your thinking through rigorous study.

There are in fact multiple scientific methods, as the basic structure can be easily modified.  The one we typically learn about in school is the basic method, based in logic and problem solving, typically used in “hard” science fields like biology, chemistry, and physics. It may vary in other fields, such as psychology, but the basic premise of making observations, testing, and continuing to improve a theory from the results remain the same.

body_history

The History of the Scientific Method

The scientific method as we know it today is based on thousands of years of scientific study. Its development goes all the way back to ancient Mesopotamia, Greece, and India.

The Ancient World

In ancient Greece, Aristotle devised an inductive-deductive process , which weighs broad generalizations from data against conclusions reached by narrowing down possibilities from a general statement. However, he favored deductive reasoning, as it identifies causes, which he saw as more important.

Aristotle wrote a great deal about logic and many of his ideas about reasoning echo those found in the modern scientific method, such as ignoring circular evidence and limiting the number of middle terms between the beginning of an experiment and the end. Though his model isn’t the one that we use today, the reliance on logic and thorough testing are still key parts of science today.

The Middle Ages

The next big step toward the development of the modern scientific method came in the Middle Ages, particularly in the Islamic world. Ibn al-Haytham, a physicist from what we now know as Iraq, developed a method of testing, observing, and deducing for his research on vision. al-Haytham was critical of Aristotle’s lack of inductive reasoning, which played an important role in his own research.

Other scientists, including Abū Rayhān al-Bīrūnī, Ibn Sina, and Robert Grosseteste also developed models of scientific reasoning to test their own theories. Though they frequently disagreed with one another and Aristotle, those disagreements and refinements of their methods led to the scientific method we have today.

Following those major developments, particularly Grosseteste’s work, Roger Bacon developed his own cycle of observation (seeing that something occurs), hypothesis (making a guess about why that thing occurs), experimentation (testing that the thing occurs), and verification (an outside person ensuring that the result of the experiment is consistent).

After joining the Franciscan Order, Bacon was granted a special commission to write about science; typically, Friars were not allowed to write books or pamphlets. With this commission, Bacon outlined important tenets of the scientific method, including causes of error, methods of knowledge, and the differences between speculative and experimental science. He also used his own principles to investigate the causes of a rainbow, demonstrating the method’s effectiveness.

Scientific Revolution

Throughout the Renaissance, more great thinkers became involved in devising a thorough, rigorous method of scientific study. Francis Bacon brought inductive reasoning further into the method, whereas Descartes argued that the laws of the universe meant that deductive reasoning was sufficient. Galileo’s research was also inductive reasoning-heavy, as he believed that researchers could not account for every possible variable; therefore, repetition was necessary to eliminate faulty hypotheses and experiments.

All of this led to the birth of the Scientific Revolution , which took place during the sixteenth and seventeenth centuries. In 1660, a group of philosophers and physicians joined together to work on scientific advancement. After approval from England’s crown , the group became known as the Royal Society, which helped create a thriving scientific community and an early academic journal to help introduce rigorous study and peer review.

Previous generations of scientists had touched on the importance of induction and deduction, but Sir Isaac Newton proposed that both were equally important. This contribution helped establish the importance of multiple kinds of reasoning, leading to more rigorous study.

As science began to splinter into separate areas of study, it became necessary to define different methods for different fields. Karl Popper was a leader in this area—he established that science could be subject to error, sometimes intentionally. This was particularly tricky for “soft” sciences like psychology and social sciences, which require different methods. Popper’s theories furthered the divide between sciences like psychology and “hard” sciences like chemistry or physics.

Paul Feyerabend argued that Popper’s methods were too restrictive for certain fields, and followed a less restrictive method hinged on “anything goes,” as great scientists had made discoveries without the Scientific Method. Feyerabend suggested that throughout history scientists had adapted their methods as necessary, and that sometimes it would be necessary to break the rules. This approach suited social and behavioral scientists particularly well, leading to a more diverse range of models for scientists in multiple fields to use.

body_experiment-3

The Scientific Method Steps

Though different fields may have variations on the model, the basic scientific method is as follows:

#1: Make Observations 

Notice something, such as the air temperature during the winter, what happens when ice cream melts, or how your plants behave when you forget to water them.

#2: Ask a Question

Turn your observation into a question. Why is the temperature lower during the winter? Why does my ice cream melt? Why does my toast always fall butter-side down?

This step can also include doing some research. You may be able to find answers to these questions already, but you can still test them!

#3: Make a Hypothesis

A hypothesis is an educated guess of the answer to your question. Why does your toast always fall butter-side down? Maybe it’s because the butter makes that side of the bread heavier.

A good hypothesis leads to a prediction that you can test, phrased as an if/then statement. In this case, we can pick something like, “If toast is buttered, then it will hit the ground butter-first.”

#4: Experiment

Your experiment is designed to test whether your predication about what will happen is true. A good experiment will test one variable at a time —for example, we’re trying to test whether butter weighs down one side of toast, making it more likely to hit the ground first.

The unbuttered toast is our control variable. If we determine the chance that a slice of unbuttered toast, marked with a dot, will hit the ground on a particular side, we can compare those results to our buttered toast to see if there’s a correlation between the presence of butter and which way the toast falls.

If we decided not to toast the bread, that would be introducing a new question—whether or not toasting the bread has any impact on how it falls. Since that’s not part of our test, we’ll stick with determining whether the presence of butter has any impact on which side hits the ground first.

#5: Analyze Data

After our experiment, we discover that both buttered toast and unbuttered toast have a 50/50 chance of hitting the ground on the buttered or marked side when dropped from a consistent height, straight down. It looks like our hypothesis was incorrect—it’s not the butter that makes the toast hit the ground in a particular way, so it must be something else.

Since we didn’t get the desired result, it’s back to the drawing board. Our hypothesis wasn’t correct, so we’ll need to start fresh. Now that you think about it, your toast seems to hit the ground butter-first when it slides off your plate, not when you drop it from a consistent height. That can be the basis for your new experiment.

#6: Communicate Your Results

Good science needs verification. Your experiment should be replicable by other people, so you can put together a report about how you ran your experiment to see if other peoples’ findings are consistent with yours.

This may be useful for class or a science fair. Professional scientists may publish their findings in scientific journals, where other scientists can read and attempt their own versions of the same experiments. Being part of a scientific community helps your experiments be stronger because other people can see if there are flaws in your approach—such as if you tested with different kinds of bread, or sometimes used peanut butter instead of butter—that can lead you closer to a good answer.

body_toast-1

A Scientific Method Example: Falling Toast

We’ve run through a quick recap of the scientific method steps, but let’s look a little deeper by trying again to figure out why toast so often falls butter side down.

#1: Make Observations

At the end of our last experiment, where we learned that butter doesn’t actually make toast more likely to hit the ground on that side, we remembered that the times when our toast hits the ground butter side first are usually when it’s falling off a plate.

The easiest question we can ask is, “Why is that?”

We can actually search this online and find a pretty detailed answer as to why this is true. But we’re budding scientists—we want to see it in action and verify it for ourselves! After all, good science should be replicable, and we have all the tools we need to test out what’s really going on.

Why do we think that buttered toast hits the ground butter-first? We know it’s not because it’s heavier, so we can strike that out. Maybe it’s because of the shape of our plate?

That’s something we can test. We’ll phrase our hypothesis as, “If my toast slides off my plate, then it will fall butter-side down.”

Just seeing that toast falls off a plate butter-side down isn’t enough for us. We want to know why, so we’re going to take things a step further—we’ll set up a slow-motion camera to capture what happens as the toast slides off the plate.

We’ll run the test ten times, each time tilting the same plate until the toast slides off. We’ll make note of each time the butter side lands first and see what’s happening on the video so we can see what’s going on.

When we review the footage, we’ll likely notice that the bread starts to flip when it slides off the edge, changing how it falls in a way that didn’t happen when we dropped it ourselves.

That answers our question, but it’s not the complete picture —how do other plates affect how often toast hits the ground butter-first? What if the toast is already butter-side down when it falls? These are things we can test in further experiments with new hypotheses!

Now that we have results, we can share them with others who can verify our results. As mentioned above, being part of the scientific community can lead to better results. If your results were wildly different from the established thinking about buttered toast, that might be cause for reevaluation. If they’re the same, they might lead others to make new discoveries about buttered toast. At the very least, you have a cool experiment you can share with your friends!

Key Scientific Method Tips

Though science can be complex, the benefit of the scientific method is that it gives you an easy-to-follow means of thinking about why and how things happen. To use it effectively, keep these things in mind!

Don’t Worry About Proving Your Hypothesis

One of the important things to remember about the scientific method is that it’s not necessarily meant to prove your hypothesis right. It’s great if you do manage to guess the reason for something right the first time, but the ultimate goal of an experiment is to find the true reason for your observation to occur, not to prove your hypothesis right.

Good science sometimes means that you’re wrong. That’s not a bad thing—a well-designed experiment with an unanticipated result can be just as revealing, if not more, than an experiment that confirms your hypothesis.

Be Prepared to Try Again

If the data from your experiment doesn’t match your hypothesis, that’s not a bad thing. You’ve eliminated one possible explanation, which brings you one step closer to discovering the truth.

The scientific method isn’t something you’re meant to do exactly once to prove a point. It’s meant to be repeated and adapted to bring you closer to a solution. Even if you can demonstrate truth in your hypothesis, a good scientist will run an experiment again to be sure that the results are replicable. You can even tweak a successful hypothesis to test another factor, such as if we redid our buttered toast experiment to find out whether different kinds of plates affect whether or not the toast falls butter-first. The more we test our hypothesis, the stronger it becomes!

What’s Next?

Want to learn more about the scientific method? These important high school science classes will no doubt cover it in a variety of different contexts.

Test your ability to follow the scientific method using these at-home science experiments for kids !

Need some proof that science is fun? Try making slime

Trending Now

How to Get Into Harvard and the Ivy League

How to Get a Perfect 4.0 GPA

How to Write an Amazing College Essay

What Exactly Are Colleges Looking For?

ACT vs. SAT: Which Test Should You Take?

When should you take the SAT or ACT?

Get Your Free

PrepScholar

Find Your Target SAT Score

Free Complete Official SAT Practice Tests

How to Get a Perfect SAT Score, by an Expert Full Scorer

Score 800 on SAT Math

Score 800 on SAT Reading and Writing

How to Improve Your Low SAT Score

Score 600 on SAT Math

Score 600 on SAT Reading and Writing

Find Your Target ACT Score

Complete Official Free ACT Practice Tests

How to Get a Perfect ACT Score, by a 36 Full Scorer

Get a 36 on ACT English

Get a 36 on ACT Math

Get a 36 on ACT Reading

Get a 36 on ACT Science

How to Improve Your Low ACT Score

Get a 24 on ACT English

Get a 24 on ACT Math

Get a 24 on ACT Reading

Get a 24 on ACT Science

Stay Informed

Get the latest articles and test prep tips!

Follow us on Facebook (icon)

Melissa Brinks graduated from the University of Washington in 2014 with a Bachelor's in English with a creative writing emphasis. She has spent several years tutoring K-12 students in many subjects, including in SAT prep, to help them prepare for their college education.

Ask a Question Below

Have any questions about this article or other topics? Ask below and we'll reply!

  • Bipolar Disorder
  • Therapy Center
  • When To See a Therapist
  • Types of Therapy
  • Best Online Therapy
  • Best Couples Therapy
  • Best Family Therapy
  • Managing Stress
  • Sleep and Dreaming
  • Understanding Emotions
  • Self-Improvement
  • Healthy Relationships
  • Student Resources
  • Personality Types
  • Sweepstakes
  • Guided Meditations
  • Verywell Mind Insights
  • 2024 Verywell Mind 25
  • Mental Health in the Classroom
  • Editorial Process
  • Meet Our Review Board
  • Crisis Support

Scientific Method Steps in Psychology Research

Steps, Uses, and Key Terms

Verywell / Theresa Chiechi

How do researchers investigate psychological phenomena? They utilize a process known as the scientific method to study different aspects of how people think and behave.

When conducting research, the scientific method steps to follow are:

  • Observe what you want to investigate
  • Ask a research question and make predictions
  • Test the hypothesis and collect data
  • Examine the results and draw conclusions
  • Report and share the results 

This process not only allows scientists to investigate and understand different psychological phenomena but also provides researchers and others a way to share and discuss the results of their studies.

Generally, there are five main steps in the scientific method, although some may break down this process into six or seven steps. An additional step in the process can also include developing new research questions based on your findings.

What Is the Scientific Method?

What is the scientific method and how is it used in psychology?

The scientific method consists of five steps. It is essentially a step-by-step process that researchers can follow to determine if there is some type of relationship between two or more variables.

By knowing the steps of the scientific method, you can better understand the process researchers go through to arrive at conclusions about human behavior.

Scientific Method Steps

While research studies can vary, these are the basic steps that psychologists and scientists use when investigating human behavior.

The following are the scientific method steps:

Step 1. Make an Observation

Before a researcher can begin, they must choose a topic to study. Once an area of interest has been chosen, the researchers must then conduct a thorough review of the existing literature on the subject. This review will provide valuable information about what has already been learned about the topic and what questions remain to be answered.

A literature review might involve looking at a considerable amount of written material from both books and academic journals dating back decades.

The relevant information collected by the researcher will be presented in the introduction section of the final published study results. This background material will also help the researcher with the first major step in conducting a psychology study: formulating a hypothesis.

Step 2. Ask a Question

Once a researcher has observed something and gained some background information on the topic, the next step is to ask a question. The researcher will form a hypothesis, which is an educated guess about the relationship between two or more variables

For example, a researcher might ask a question about the relationship between sleep and academic performance: Do students who get more sleep perform better on tests at school?

In order to formulate a good hypothesis, it is important to think about different questions you might have about a particular topic.

You should also consider how you could investigate the causes. Falsifiability is an important part of any valid hypothesis. In other words, if a hypothesis was false, there needs to be a way for scientists to demonstrate that it is false.

Step 3. Test Your Hypothesis and Collect Data

Once you have a solid hypothesis, the next step of the scientific method is to put this hunch to the test by collecting data. The exact methods used to investigate a hypothesis depend on exactly what is being studied. There are two basic forms of research that a psychologist might utilize: descriptive research or experimental research.

Descriptive research is typically used when it would be difficult or even impossible to manipulate the variables in question. Examples of descriptive research include case studies, naturalistic observation , and correlation studies. Phone surveys that are often used by marketers are one example of descriptive research.

Correlational studies are quite common in psychology research. While they do not allow researchers to determine cause-and-effect, they do make it possible to spot relationships between different variables and to measure the strength of those relationships. 

Experimental research is used to explore cause-and-effect relationships between two or more variables. This type of research involves systematically manipulating an independent variable and then measuring the effect that it has on a defined dependent variable .

One of the major advantages of this method is that it allows researchers to actually determine if changes in one variable actually cause changes in another.

While psychology experiments are often quite complex, a simple experiment is fairly basic but does allow researchers to determine cause-and-effect relationships between variables. Most simple experiments use a control group (those who do not receive the treatment) and an experimental group (those who do receive the treatment).

Step 4. Examine the Results and Draw Conclusions

Once a researcher has designed the study and collected the data, it is time to examine this information and draw conclusions about what has been found.  Using statistics , researchers can summarize the data, analyze the results, and draw conclusions based on this evidence.

So how does a researcher decide what the results of a study mean? Not only can statistical analysis support (or refute) the researcher’s hypothesis; it can also be used to determine if the findings are statistically significant.

When results are said to be statistically significant, it means that it is unlikely that these results are due to chance.

Based on these observations, researchers must then determine what the results mean. In some cases, an experiment will support a hypothesis, but in other cases, it will fail to support the hypothesis.

So what happens if the results of a psychology experiment do not support the researcher's hypothesis? Does this mean that the study was worthless?

Just because the findings fail to support the hypothesis does not mean that the research is not useful or informative. In fact, such research plays an important role in helping scientists develop new questions and hypotheses to explore in the future.

After conclusions have been drawn, the next step is to share the results with the rest of the scientific community. This is an important part of the process because it contributes to the overall knowledge base and can help other scientists find new research avenues to explore.

Step 5. Report the Results

The final step in a psychology study is to report the findings. This is often done by writing up a description of the study and publishing the article in an academic or professional journal. The results of psychological studies can be seen in peer-reviewed journals such as  Psychological Bulletin , the  Journal of Social Psychology ,  Developmental Psychology , and many others.

The structure of a journal article follows a specified format that has been outlined by the  American Psychological Association (APA) . In these articles, researchers:

  • Provide a brief history and background on previous research
  • Present their hypothesis
  • Identify who participated in the study and how they were selected
  • Provide operational definitions for each variable
  • Describe the measures and procedures that were used to collect data
  • Explain how the information collected was analyzed
  • Discuss what the results mean

Why is such a detailed record of a psychological study so important? By clearly explaining the steps and procedures used throughout the study, other researchers can then replicate the results. The editorial process employed by academic and professional journals ensures that each article that is submitted undergoes a thorough peer review, which helps ensure that the study is scientifically sound.

Once published, the study becomes another piece of the existing puzzle of our knowledge base on that topic.

Before you begin exploring the scientific method steps, here's a review of some key terms and definitions that you should be familiar with:

  • Falsifiable : The variables can be measured so that if a hypothesis is false, it can be proven false
  • Hypothesis : An educated guess about the possible relationship between two or more variables
  • Variable : A factor or element that can change in observable and measurable ways
  • Operational definition : A full description of exactly how variables are defined, how they will be manipulated, and how they will be measured

Uses for the Scientific Method

The  goals of psychological studies  are to describe, explain, predict and perhaps influence mental processes or behaviors. In order to do this, psychologists utilize the scientific method to conduct psychological research. The scientific method is a set of principles and procedures that are used by researchers to develop questions, collect data, and reach conclusions.

Goals of Scientific Research in Psychology

Researchers seek not only to describe behaviors and explain why these behaviors occur; they also strive to create research that can be used to predict and even change human behavior.

Psychologists and other social scientists regularly propose explanations for human behavior. On a more informal level, people make judgments about the intentions, motivations , and actions of others on a daily basis.

While the everyday judgments we make about human behavior are subjective and anecdotal, researchers use the scientific method to study psychology in an objective and systematic way. The results of these studies are often reported in popular media, which leads many to wonder just how or why researchers arrived at the conclusions they did.

Examples of the Scientific Method

Now that you're familiar with the scientific method steps, it's useful to see how each step could work with a real-life example.

Say, for instance, that researchers set out to discover what the relationship is between psychotherapy and anxiety .

  • Step 1. Make an observation : The researchers choose to focus their study on adults ages 25 to 40 with generalized anxiety disorder.
  • Step 2. Ask a question : The question they want to answer in their study is: Do weekly psychotherapy sessions reduce symptoms in adults ages 25 to 40 with generalized anxiety disorder?
  • Step 3. Test your hypothesis : Researchers collect data on participants' anxiety symptoms . They work with therapists to create a consistent program that all participants undergo. Group 1 may attend therapy once per week, whereas group 2 does not attend therapy.
  • Step 4. Examine the results : Participants record their symptoms and any changes over a period of three months. After this period, people in group 1 report significant improvements in their anxiety symptoms, whereas those in group 2 report no significant changes.
  • Step 5. Report the results : Researchers write a report that includes their hypothesis, information on participants, variables, procedure, and conclusions drawn from the study. In this case, they say that "Weekly therapy sessions are shown to reduce anxiety symptoms in adults ages 25 to 40."

Of course, there are many details that go into planning and executing a study such as this. But this general outline gives you an idea of how an idea is formulated and tested, and how researchers arrive at results using the scientific method.

Erol A. How to conduct scientific research ? Noro Psikiyatr Ars . 2017;54(2):97-98. doi:10.5152/npa.2017.0120102

University of Minnesota. Psychologists use the scientific method to guide their research .

Shaughnessy, JJ, Zechmeister, EB, & Zechmeister, JS. Research Methods In Psychology . New York: McGraw Hill Education; 2015.

By Kendra Cherry, MSEd Kendra Cherry, MS, is a psychosocial rehabilitation specialist, psychology educator, and author of the "Everything Psychology Book."

  • A to Z Guides

What Is the Scientific Method?

characteristics of scientific method of problem solving

The scientific method is a systematic way of conducting experiments or studies so that you can explore the things you observe in the world and answer questions about them. The scientific method, also known as the hypothetico-deductive method, is a series of steps that can help you accurately describe the things you observe or improve your understanding of them.

Ultimately, your goal when you use the scientific method is to:

  • Find a cause-and-effect relationship by asking a question about something you observed
  • Collect as much evidence as you can about what you observed, as this can help you explore the connection between your evidence and what you observed
  • Determine if all your evidence can be combined to answer your question in a way that makes sense

Francis Bacon and René Descartes are usually credited with formalizing the process in the 16th and 17th centuries. The two philosophers argued that research shouldn’t be guided by preset metaphysical ideas of how reality works. They supported the use of inductive reasoning to come up with hypotheses and understand new things about reality.

Scientific Method Steps

The scientific method is a step-by-step problem-solving process. These steps include:

Observe the world around you. This will help you come up with a topic you are interested in and want to learn more about. In many cases, you already have a topic in mind because you have a related question for which you couldn't find an immediate answer.

Either way, you'll start the process by finding out what people before you already know about the topic, as well as any questions that people are still asking about. You may need to look up and read books and articles from academic journals or talk to other people so that you understand as much as you possibly can about your topic. This will help you with your next step.

Ask questions. Asking questions about what you observed and learned from reading and talking to others can help you figure out what the "problem" is. Scientists try to ask questions that are both interesting and specific and can be answered with the help of a fairly easy experiment or series of experiments. Your question should have one part (called a variable) that you can change in your experiment and another variable that you can measure. Your goal is to design an experiment that is a "fair test," which is when all the conditions in the experiment are kept the same except for the one you change (called the experimental or independent variable).

Form a hypothesis and make predictions based on it.  A hypothesis is an educated guess about the relationship between two or more variables in your question. A good hypothesis lets you predict what will happen when you test it in an experiment. Another important feature of a good hypothesis is that, if the hypothesis is wrong, you should be able to show that it's wrong. This is called falsifiability. If your experiment shows that your prediction is true, then your hypothesis is supported by your data.

Test your prediction by doing an experiment or making more observations.  The way you test your prediction depends on what you are studying. The best support comes from an experiment, but in some cases, it's too hard or impossible to change the variables in an experiment. Sometimes, you may need to do descriptive research where you gather more observations instead of doing an experiment. You will carefully gather notes and measurements during your experiments or studies, and you can share them with other people interested in the same question as you. Ideally, you will also repeat your experiment a couple more times because it's possible to get a result by chance, but it's less possible to get the same result more than once by chance.

Draw a conclusion. You will analyze what you already know about your topic from your literature research and the data gathered during your experiment. This will help you decide if the conclusion you draw from your data supports or contradicts your hypothesis. If your results contradict your hypothesis, you can use this observation to form a new hypothesis and make a new prediction. This is why scientific research is ongoing and scientific knowledge is changing all the time. It's very common for scientists to get results that don't support their hypotheses. In fact, you sometimes learn more about the world when your experiments don't support your hypotheses because it leads you to ask more questions. And this time around, you already know that one possible explanation is likely wrong.

Use your results to guide your next steps (iterate). For instance, if your hypothesis is supported, you may do more experiments to confirm it. Or you could come up with a hypothesis about why it works this way and design an experiment to test that. If your hypothesis is not supported, you can come up with another hypothesis and do experiments to test it. You'll rarely get the right hypothesis in one go. Most of the time, you'll have to go back to the hypothesis stage and try again. Every attempt offers you important information that helps you improve your next round of questions, hypotheses, and predictions.

Share your results. Scientific research isn't something you can do on your own; you must work with other people to do it.   You may be able to do an experiment or a series of experiments on your own, but you can't come up with all the ideas or do all the experiments by yourself .

Scientists and researchers usually share information by publishing it in a scientific journal or by presenting it to their colleagues during meetings and scientific conferences. These journals are read and the conferences are attended by other researchers who are interested in the same questions. If there's anything wrong with your hypothesis, prediction, experiment design, or conclusion, other researchers will likely find it and point it out to you.

It can be scary, but it's a critical part of doing scientific research. You must let your research be examined by other researchers who are as interested and knowledgeable about your question as you. This process helps other researchers by pointing out hypotheses that have been proved wrong and why they are wrong. It helps you by identifying flaws in your thinking or experiment design. And if you don't share what you've learned and let other people ask questions about it, it's not helpful to your or anyone else's understanding of what happens in the world.

Scientific Method Example

Here's an everyday example of how you can apply the scientific method to understand more about your world so you can solve your problems in a helpful way.

Let's say you put slices of bread in your toaster and press the button, but nothing happens. Your toaster isn't working, but you can't afford to buy a new one right now. You might be able to rescue it from the trash can if you can figure out what's wrong with it. So, let's figure out what's wrong with your toaster.

Observation. Your toaster isn't working to toast your bread.

Ask a question. In this case, you're asking, "Why isn't my toaster working?" You could even do a bit of preliminary research by looking in the owner's manual for your toaster. The manufacturer has likely tested your toaster model under many conditions, and they may have some ideas for where to start with your hypothesis.

Form a hypothesis and make predictions based on it. Your hypothesis should be a potential explanation or answer to the question that you can test to see if it's correct. One possible explanation that we could test is that the power outlet is broken. Our prediction is that if the outlet is broken, then plugging it into a different outlet should make the toaster work again.

Test your prediction by doing an experiment or making more observations. You plug the toaster into a different outlet and try to toast your bread.

If that works, then your hypothesis is supported by your experimental data. Results that support your hypothesis don't prove it right; they simply suggest that it's a likely explanation. This uncertainty arises because, in the real world, we can't rule out the possibility of mistakes, wrong assumptions, or weird coincidences affecting the results. If the toaster doesn’t work even after plugging it into a different outlet, then your hypothesis is not supported and it's likely the wrong explanation.

Use your results to guide your next steps (iteration). If your toaster worked, you may decide to do further tests to confirm it or revise it. For example, you could plug something else that you know is working into the first outlet to see if that stops working too. That would be further confirmation that your hypothesis is correct.

If your toaster failed to toast when plugged into the second outlet, you need a new hypothesis. For example, your next hypothesis might be that the toaster has a shorted wire. You could test this hypothesis directly if you have the right equipment and training, or you could take it to a repair shop where they could test that hypothesis for you.

Share your results. For this everyday example, you probably wouldn't want to write a paper, but you could share your problem-solving efforts with your housemates or anyone you hire to repair your outlet or help you test if the toaster has a short circuit.

What the Scientific Method Is Used For

The scientific method is useful whenever you need to reason logically about your questions and gather evidence to support your problem-solving efforts. So, you can use it in everyday life to answer many of your questions; however, when most people think of the scientific method, they likely think of using it to:

Describe how nature works . It can be hard to accurately describe how nature works because it's almost impossible to account for every variable that's involved in a natural process. Researchers may not even know about many of the variables that are involved. In some cases, all you can do is make assumptions. But you can use the scientific method to logically disprove wrong assumptions by identifying flaws in the reasoning.

Do scientific research in a laboratory to develop things such as new medicines.

Develop critical thinking skills.  Using the scientific method may help you develop critical thinking in your daily life because you learn to systematically ask questions and gather evidence to find answers. Without logical reasoning, you might be more likely to have a distorted perspective or bias. Bias is the inclination we all have to favor one perspective (usually our own) over another.

The scientific method doesn't perfectly solve the problem of bias, but it does make it harder for an entire field to be biased in the same direction. That's because it's unlikely that all the people working in a field have the same biases. It also helps make the biases of individuals more obvious because if you repeatedly misinterpret information in the same way in multiple experiments or over a period, the other people working on the same question will notice. If you don't correct your bias when others point it out to you, you'll lose your credibility. Other people might then stop believing what you have to say.

Why Is the Scientific Method Important?

When you use the scientific method, your goal is to do research in a fair, unbiased, and repeatable way. The scientific method helps meet these goals because:

It's a systematic approach to problem-solving. It can help you figure out where you're going wrong in your thinking and research if you're not getting helpful answers to your questions. Helpful answers solve problems and keep you moving forward. So, a systematic approach helps you improve your problem-solving abilities if you get stuck.

It can help you solve your problems.  The scientific method helps you isolate problems by focusing on what's important. In addition, it can help you make your solutions better every time you go through the process.

It helps you eliminate (or become aware of) your personal biases.  It can help you limit the influence of your own personal, preconceived notions . A big part of the process is considering what other people already know and think about your question. It also involves sharing what you've learned and letting other people ask about your methods and conclusions. At the end of the process, even if you still think your answer is best, you have considered what other people know and think about the question.

The scientific method is a systematic way of conducting experiments or studies so that you can explore the world around you and answer questions using reason and evidence. It's a step-by-step problem-solving process that involves: (1) observation, (2) asking questions, (3) forming hypotheses and making predictions, (4) testing your hypotheses through experiments or more observations, (5) using what you learned through experiment or observation to guide further investigation, and (6) sharing your results.

Top doctors in ,

Find more top doctors on, related links.

  • Health A-Z News
  • Health A-Z Reference
  • Health A-Z Slideshows
  • Health A-Z Quizzes
  • Health A-Z Videos
  • WebMDRx Savings Card
  • Coronavirus (COVID-19)
  • Hepatitis C
  • Diabetes Warning Signs
  • Rheumatoid Arthritis
  • Morning-After Pill
  • Breast Cancer Screening
  • Psoriatic Arthritis Symptoms
  • Heart Failure
  • Multiple Myeloma
  • Types of Crohn's Disease

characteristics of scientific method of problem solving

Sciencing_Icons_Science SCIENCE

Sciencing_icons_biology biology, sciencing_icons_cells cells, sciencing_icons_molecular molecular, sciencing_icons_microorganisms microorganisms, sciencing_icons_genetics genetics, sciencing_icons_human body human body, sciencing_icons_ecology ecology, sciencing_icons_chemistry chemistry, sciencing_icons_atomic &amp; molecular structure atomic & molecular structure, sciencing_icons_bonds bonds, sciencing_icons_reactions reactions, sciencing_icons_stoichiometry stoichiometry, sciencing_icons_solutions solutions, sciencing_icons_acids &amp; bases acids & bases, sciencing_icons_thermodynamics thermodynamics, sciencing_icons_organic chemistry organic chemistry, sciencing_icons_physics physics, sciencing_icons_fundamentals-physics fundamentals, sciencing_icons_electronics electronics, sciencing_icons_waves waves, sciencing_icons_energy energy, sciencing_icons_fluid fluid, sciencing_icons_astronomy astronomy, sciencing_icons_geology geology, sciencing_icons_fundamentals-geology fundamentals, sciencing_icons_minerals &amp; rocks minerals & rocks, sciencing_icons_earth scructure earth structure, sciencing_icons_fossils fossils, sciencing_icons_natural disasters natural disasters, sciencing_icons_nature nature, sciencing_icons_ecosystems ecosystems, sciencing_icons_environment environment, sciencing_icons_insects insects, sciencing_icons_plants &amp; mushrooms plants & mushrooms, sciencing_icons_animals animals, sciencing_icons_math math, sciencing_icons_arithmetic arithmetic, sciencing_icons_addition &amp; subtraction addition & subtraction, sciencing_icons_multiplication &amp; division multiplication & division, sciencing_icons_decimals decimals, sciencing_icons_fractions fractions, sciencing_icons_conversions conversions, sciencing_icons_algebra algebra, sciencing_icons_working with units working with units, sciencing_icons_equations &amp; expressions equations & expressions, sciencing_icons_ratios &amp; proportions ratios & proportions, sciencing_icons_inequalities inequalities, sciencing_icons_exponents &amp; logarithms exponents & logarithms, sciencing_icons_factorization factorization, sciencing_icons_functions functions, sciencing_icons_linear equations linear equations, sciencing_icons_graphs graphs, sciencing_icons_quadratics quadratics, sciencing_icons_polynomials polynomials, sciencing_icons_geometry geometry, sciencing_icons_fundamentals-geometry fundamentals, sciencing_icons_cartesian cartesian, sciencing_icons_circles circles, sciencing_icons_solids solids, sciencing_icons_trigonometry trigonometry, sciencing_icons_probability-statistics probability & statistics, sciencing_icons_mean-median-mode mean/median/mode, sciencing_icons_independent-dependent variables independent/dependent variables, sciencing_icons_deviation deviation, sciencing_icons_correlation correlation, sciencing_icons_sampling sampling, sciencing_icons_distributions distributions, sciencing_icons_probability probability, sciencing_icons_calculus calculus, sciencing_icons_differentiation-integration differentiation/integration, sciencing_icons_application application, sciencing_icons_projects projects, sciencing_icons_news news.

  • Share Tweet Email Print
  • Home ⋅
  • Science Fair Project Ideas for Kids, Middle & High School Students ⋅
  • Probability & Statistics

Five Characteristics of the Scientific Method

Five Characteristics of the Scientific Method

Steps & Procedures for Conducting Scientific Research

The scientific method is the system used by scientists to explore data, generate and test hypotheses, develop new theories and confirm or reject earlier results. Although the exact methods used in the different sciences vary (for example, physicists and psychologists work in very different ways), they share some fundamental attributes that may be called characteristics of the scientific method.

TL;DR (Too Long; Didn't Read)

Five key descriptors for the scientific method are: empirical, replicable, provisional, objective and systematic.

Empirical Observation

The scientific method is empirical. That is, it relies on direct observation of the world, and disdains hypotheses that run counter to observable fact. This contrasts with methods that rely on pure reason (including that proposed by Plato) and with methods that rely on emotional or other subjective factors.

Replicable Experiments

Scientific experiments are replicable. That is, if another person duplicates the experiment, he or she will get the same results. Scientists are supposed to publish enough of their method so that another person, with appropriate training, could replicate the results. This contrasts with methods that rely on experiences that are unique to a particular individual or a small group of individuals.

Provisional Results

Results obtained through the scientific method are provisional; they are (or ought to be) open to question and debate. If new data arise that contradict a theory, that theory must be modified. For example, the phlogiston theory of fire and combustion was rejected when evidence against it arose.

Objective Approach

The scientific method is objective. It relies on facts and on the world as it is, rather than on beliefs, wishes or desires. Scientists attempt (with varying degrees of success) to remove their biases when making observations.

Systematic Observation

Strictly speaking, the scientific method is systematic; that is, it relies on carefully planned studies rather than on random or haphazard observation. Nevertheless, science can begin from some random observation. Isaac Asimov said that the most exciting phrase to hear in science is not "Eureka!" but "That's funny." After the scientist notices something funny, he or she proceeds to investigate it systematically.

Related Articles

Research methods in science, difference between proposition & hypothesis, 10 characteristics of a science experiment, how to eliminate bias in qualitative research, characteristics of modern science, what are the 8 steps in scientific research, differences between conceptual independent variables..., what are main limitations of behavioral theories, types of observation in the scientific method, the differences between concepts, theories & paradigms, the importance of hypothesis testing, the definition of an uncontrolled variable, 5 components of a well-designed scientific experiment, what is the purpose of factor analysis, distinguishing between descriptive & causal studies, how to write a summary on a science project, what is the next step if an experiment fails to confirm..., scientists now know why you sometimes feel psychic.

  • Scientific Method

About the Author

Peter Flom is a statistician and a learning-disabled adult. He has been writing for many years and has been published in many academic journals in fields such as psychology, drug addiction, epidemiology and others. He holds a Ph.D. in psychometrics from Fordham University.

Find Your Next Great Science Fair Project! GO

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

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

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

Biology archive

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

  • Controlled experiments
  • The scientific method and experimental design

characteristics of scientific method of problem solving

Introduction

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

Scientific method example: Failure to toast

1. make an observation., 2. ask a question., 3. propose a hypothesis., 4. make predictions., 5. test the predictions..

  • If the toaster does toast, then the hypothesis is supported—likely correct.
  • If the toaster doesn't toast, then the hypothesis is not supported—likely wrong.

Logical possibility

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

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

Want to join the conversation?

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

Incredible Answer

medieval alchemist working in his lab with flasks in foreground

Multiple goals, multiple solutions, plenty of second-guessing and revising − here’s how science really works

characteristics of scientific method of problem solving

Professor of Philosophy, University of Montana

Disclosure statement

Soazig Le Bihan receives funding from the Maureen and Mike Mansfield Center at the University of Montana.

University of Montana provides funding as a member of The Conversation US.

View all partners

A man in a lab coat bends under a dim light, his strained eyes riveted onto a microscope. He’s powered only by caffeine and anticipation.

This solitary scientist will stay on task until he unveils the truth about the cause of the dangerous disease quickly spreading through his vulnerable city. Time is short, the stakes are high, and only he can save everyone. …

That kind of romanticized picture of science was standard for a long time. But it’s as far from actual scientific practice as a movie’s choreographed martial arts battle is from a real fistfight.

For most of the 20th century, philosophers of science like me maintained somewhat idealistic claims about what good science looks like. Over the past few decades, however, many of us have revised our views to better mirror actual scientific practice .

An update on what to expect from actual science is overdue. I often worry that when the public holds science to unrealistic standards, any scientific claim failing to live up to them arouses suspicion. While public trust is globally strong and has been for decades, it has been eroding. In November 2023, Americans’ trust in scientists was 14 points lower than it had been just prior to the COVID-19 pandemic, with its flurry of confusing and sometimes contradictory science-related messages.

When people’s expectations are not met about how science works, they may blame scientists. But modifying our expectations might be more useful. Here are three updates I think can help people better understand how science actually works. Hopefully, a better understanding of actual scientific practice will also shore up people’s trust in the process.

The many faces of scientific research

First, science is a complex endeavor involving multiple goals and associated activities.

Some scientists search for the causes underlying some observable effect, such as a decimated pine forest or the Earth’s global surface temperature increase .

Others may investigate the what rather than the why of things. For example, ecologists build models to estimate gray wolf abundance in Montana . Spotting predators is incredibly challenging. Counting all of them is impractical. Abundance models are neither complete nor 100% accurate – they offer estimates deemed good enough to set harvesting quotas. Perfect scientific models are just not in the cards .

older woman holding pill bottle, medical worker in scrubs faces her

Beyond the what and the why, scientists may focus on the how. For instance, the lives of people living with chronic illnesses can be improved by research on strategies for managing disease – to mitigate symptoms and improve function, even if the true causes of their disorders largely elude current medicine.

It’s understandable that some patients may grow frustrated or distrustful of medical providers unable to give clear answers about what causes their ailment. But it’s important to grasp that lots of scientific research focuses on how to effectively intervene in the world to reach some specific goals.

Simplistic views represent science as solely focused on providing causal explanations for the various phenomena we observe in this world. The truth is that scientists tackle all kinds of problems, which are best solved using different strategies and approaches and only sometimes involve full-fledged explanations.

Complex problems call for complex solutions

The second aspect of scientific practice worth underscoring is that, because scientists tackle complex problems, they don’t typically offer one unique, complete and perfect answer. Instead they consider multiple, partial and possibly conflicting solutions.

Scientific modeling strategies illustrate this point well. Scientific models typically are partial, simplified and sometimes deliberately unrealistic representations of a system of interest. Models can be physical, conceptual or mathematical. The critical point is that they represent target systems in ways that are useful in particular contexts of inquiry. Interestingly, considering multiple possible models is often the best strategy to tackle complex problems.

Scientists consider multiple models of biodiversity , atomic nuclei or climate change . Returning to wolf abundance estimates, multiple models can also fit the bill. Such models rely on various types of data, including acoustic surveys of wolf howls, genetic methods that use fecal samples from wolves, wolf sightings and photographic evidence, aerial surveys, snow track surveys and more.

Weighing the pros and cons of various possible solutions to the problem of interest is part and parcel of the scientific process. Interestingly, in some cases, using multiple conflicting models allows for better predictions than trying to unify all the models into one.

The public may be surprised and possibly suspicious when scientists push forward multiple models that rely on conflicting assumptions and make different predictions. People often think “real science” should provide definite, complete and foolproof answers to their questions. But given various limitations and the world’s complexity, keeping multiple perspectives in play is most often the best way for scientists to reach their goals and solve the problems at hand.

woman at podium with slides beside her, presenting to a room

Science as a collective, contrarian endeavor

Finally, science is a collective endeavor, where healthy disagreement is a feature, not a bug.

The romanticized version of science pictures scientists working in isolation and establishing absolute truths. Instead, science is a social and contrarian process in which the community’s scrutiny ensures we have the best available knowledge. “Best available” does not mean “definitive,” but the best we have until we find out how to improve it. Science almost always allows for disagreements among experts.

Controversies are core to how science works at its best and are as old as Western science itself. In the 1600s, Descartes and Leibniz fought over how to best characterize the laws of dynamics and the nature of motion.

The long history of atomism provides a valuable perspective on how science is an intricate and winding process rather than a fast-delivery system of results set in stone. As Jean Baptiste Perrin conducted his 1908 experiments that seemingly settled all discussion regarding the existence of atoms and molecules, the questions of the atom’s properties were about to become the topic of decades of controversies with the birth of quantum physics.

The nature and structure of fundamental particles and associated fields have been the subject of scientific research for more than a century. Lively academic discussions abound concerning the difficult interpretation of quantum mechanics , the challenging unification of quantum physics and relativity , and the existence of the Higgs boson , among others.

Distrusting researchers for having healthy scientific disagreements is largely misguided.

A very human practice

To be clear, science is dysfunctional in some respects and contexts. Current institutions have incentives for counterproductive practices, including maximizing publication numbers . Like any human endeavor, science includes people with bad intent, including some trying to discredit legitimate scientific research . Finally, science is sometimes inappropriately influenced by various values in problematic ways.

These are all important considerations when evaluating the trustworthiness of particular scientific claims and recommendations. However, it is unfair, sometimes dangerous, to mistrust science for doing what it does at its best. Science is a multifaceted endeavor focused on solving complex problems that typically just don’t have simple solutions. Communities of experts scrutinize those solutions in hopes of providing the best available approach to tackling the problems of interest.

Science is also a fallible and collective process. Ignoring the realities of that process and holding science up to unrealistic standards may result in the public calling science out and losing trust in its reliability for the wrong reasons.

  • Philosophy of science
  • Scientific research
  • Science myths
  • Trust in science

characteristics of scientific method of problem solving

Service Delivery Consultant

characteristics of scientific method of problem solving

Newsletter and Deputy Social Media Producer

characteristics of scientific method of problem solving

College Director and Principal | Curtin College

characteristics of scientific method of problem solving

Head of School: Engineering, Computer and Mathematical Sciences

characteristics of scientific method of problem solving

Educational Designer

Scientific Methods

What is scientific method.

The Scientific method is a process with the help of which scientists try to investigate, verify, or construct an accurate and reliable version of any natural phenomena. They are done by creating an objective framework for the purpose of scientific inquiry and analysing the results scientifically to come to a conclusion that either supports or contradicts the observation made at the beginning.

Scientific Method Steps

The aim of all scientific methods is the same, that is, to analyse the observation made at the beginning. Still, various steps are adopted per the requirement of any given observation. However, there is a generally accepted sequence of steps in scientific methods.

Scientific Method

  • Observation and formulation of a question:  This is the first step of a scientific method. To start one, an observation has to be made into any observable aspect or phenomena of the universe, and a question needs to be asked about that aspect. For example, you can ask, “Why is the sky black at night? or “Why is air invisible?”
  • Data Collection and Hypothesis:  The next step involved in the scientific method is to collect all related data and formulate a hypothesis based on the observation. The hypothesis could be the cause of the phenomena, its effect, or its relation to any other phenomena.
  • Testing the hypothesis:  After the hypothesis is made, it needs to be tested scientifically. Scientists do this by conducting experiments. The aim of these experiments is to determine whether the hypothesis agrees with or contradicts the observations made in the real world. The confidence in the hypothesis increases or decreases based on the result of the experiments.
  • Analysis and Conclusion:  This step involves the use of proper mathematical and other scientific procedures to determine the results of the experiment. Based on the analysis, the future course of action can be determined. If the data found in the analysis is consistent with the hypothesis, it is accepted. If not, then it is rejected or modified and analysed again.

It must be remembered that a hypothesis cannot be proved or disproved by doing one experiment. It needs to be done repeatedly until there are no discrepancies in the data and the result. When there are no discrepancies and the hypothesis is proved, it is accepted as a ‘theory’.

Scientific Method Examples

Following is an example of the scientific method:

Growing bean plants:

  • What is the purpose: The main purpose of this experiment is to know where the bean plant should be kept inside or outside to check the growth rate and also set the time frame as four weeks.
  • Construction of hypothesis: The hypothesis used is that the bean plant can grow anywhere if the scientific methods are used.
  • Executing the hypothesis and collecting the data: Four bean plants are planted in identical pots using the same soil. Two are placed inside, and the other two are placed outside. Parameters like the amount of exposure to sunlight, and amount of water all are the same. After the completion of four weeks, all four plant sizes are measured.
  • Analyse the data:  While analysing the data, the average height of plants should be taken into account from both places to determine which environment is more suitable for growing the bean plants.
  • Conclusion:  The conclusion is drawn after analyzing the data.
  • Results:  Results can be reported in the form of a tabular form.

Frequently Asked Questions – FAQs

What is scientific method, what is hypothesis, give an example of a simple hypothesis., define complex hypothesis., what are the steps of the scientific method, what is the aim of scientific methods, state true or false: observation and formulation of a question is the third step of scientific method, explain the step: analysis and conclusion..

PHYSICS Related Links

Leave a Comment Cancel reply

Your Mobile number and Email id will not be published. Required fields are marked *

Request OTP on Voice Call

Post My Comment

characteristics of scientific method of problem solving

Register with BYJU'S & Download Free PDFs

Register with byju's & watch live videos.

Scientific Method characteristics, Steps-B.ed Notes

The scientific method may otherwise be called the problem-solving method. Also known as the method of science or method of scientist. The scientific method is the procedure that scientists use in the pursuit of science.

The scientific method consists of systematic observation, classification, and interpretation of data

According to M.C  Guigan- “Scientific method is a serial process by which all sciences attain their answer to their question”

For the continuous appraisal of this method, the teacher should provide such situations and activities that are conducive to its development and training.

Example: 1. the whole class can be set for any study 2. Individual laboratory experiments which involve some aspects of the scientific method may be assigned by a teacher

Characteristics of Scientific method

A method for being called a scientific method must have the following essential features-

  • Objectivity: Scientific method is quite objective in its approach and altogether free from biases and prejudices.
  • Definiteness: It is characterized by definiteness in its process as well as product. The result arrived at through the study made by this method is quite reliable and valid
  • Verifiability: Here results are not accepted unless they are verified through adequate tests and experiments.
  • Generality: The conclusion or results derived from the scientific method shows a marked characteristic of generality. Firstly, it means that the inductive method is used in making generalizations out of particular events or happenings, and secondly, the principles and laws established through this method are quite universal, having generalized application in similar other situations.
  • Predictability: In a given situation under known circumstances, what would happen to a particular object or phenomenon, can be safely predicted through the proper generalized results of the scientific method.
  • Modifiability and Dynamicity: The results obtained by the scientific method are never final, absolute, and static. They are open to verification and experimentation. Thus what is true today may prove to be wrong tomorrow on the basis of new information and findings.

Steps of the Scientific method

The steps of scientific methods are as follows-

  • Sensing the problem: A situation should be provided to the students in which they feel the need of asking and enquiring the teacher. The teacher can also raise a problem by providing such situations which stimulate reflective thinking and setting up of arriving at a rational solution. The time, availability of the material relevant to the problem, and its practical value should also be considered
  • Defining the problem: The students define their problem in scientific language and proceed towards a solution. This defining of the problem serves the ‘what’ part of their question, while the how and ‘why’ parts are yet to be in question. The teacher should help the students in framing a statement of the problem as a student in framing statement of problem a student may find it difficult to define the problem themselves
  • Analysis: The students now find the keywords and phrases in the problem which provide clues to the further study of the problem at the same time, the students must have knowledge of every keyword and an understanding of the whole problem. In our selected problem, heartbeat is the key word that gives us information.
  • Collection of data: After analyzing the problem, the teacher suggests references to the problems. The students need to plan the subsequent activities. They must discuss, consult references, and use audio-visual aids such as models, pictures, etc. while collecting data, as far as possible mechanical and personal errors should be avoided unnecessary data should also be discarded.
  • Interpreting the data: This step involves reflective thinking. This phase of problem-solving demands a great amount of guidance from the teacher because students may not be able to interpret data in the correct way due to a lack of experience. The data is organized based on similarities and differences. They can construct tables and graphs. The superfluous data should be discarded.
  • Formation of Hypothesis: After the interpretation of data, the students are asked to formulate a tentative hypothesis. A hypothesis is a probable solution to the problem at hand. The hypothesis should be free from bias and self-inclination.
  • Selection and testing of the most appropriate hypothesis: The students can select the most tenable hypothesis by rejecting others through experimentation and discussion.
  • Drawing conclusions and making generalizations: In this step, the conclusion is drawn from the selected hypothesis. The results should support the expected solution. Experiments can be repeated to verify the consistency and correctness of the conclusion drawn. A drawn conclusion should be properly reported. When some conclusions are drawn from different sets of experimentation under similar situations, they may go for the generalization of their conclusion.
  • Application of generalization to a new situation: The students should be able to apply a generalization to new situations in their daily life and hence, minimizing the gap between classroom situations and real-life situations

Scientific Method characteristics, Steps-B.ed Notes

Conclusion: Thus, the scientific method involves a definite and set procedure of attacking a problem, finding out its solution inductively, and lastly testing its adequacy of generalization by a deductive approach

Einstein Contribution to Science

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • View all journals
  • Explore content
  • About the journal
  • Publish with us
  • Sign up for alerts
  • Open access
  • Published: 10 August 2024

Digital modelling method of coal-mine roadway based on millimeter-wave radar array

  • Xusheng Xue 1 , 2 ,
  • Xingyun Yang 1 , 2 ,
  • Jianing Yue 1 , 2 ,
  • Qinghua Mao 1 , 2 ,
  • Yihan Qin 1 , 2 ,
  • Hongwei Ma 1 , 2 ,
  • Jianxin Yang 1 , 2 ,
  • Huahao Wan 1 , 2 ,
  • Enqiao Zhang 1 , 2 ,
  • Junbiao Qiu 1 , 2 ,
  • Xiaopeng Li 3 &
  • Rongquan Wang 4  

Scientific Reports volume  14 , Article number:  18585 ( 2024 ) Cite this article

Metrics details

  • Imaging techniques
  • Mechanical engineering

The roadway space of coal mine is narrow, and the illumination is low and uneven. Dynamic mining is accompanied by dust and water mist. The accuracy and reliability of roadway data collected by vision and laser sensors are poor. Based on this, a digital modeling method of coal mine roadway based on millimeter-wave radar array is proposed. Firstly, aiming at the problem of complex environmental interference, combined with the characteristics of small amount of single frame data of millimeter-wave point cloud, a multi-layer filtering noise reduction processing and dynamic subgraph registration method of millimeter-wave point cloud is proposed to filter out interference points and realize single radar point cloud registration. Secondly, aiming at the problem that a single millimeter-wave radar cannot scan the complete roadway information at one time, combined with the characteristics of small elevation field of view of millimeter-wave radar, a millimeter-wave radar array acquisition system is built, and an improved iterative closest point (ICP) registration algorithm based on point cloud features is established to construct the roadway point cloud fusion model. Finally, aiming at the problem of uneven and sparse point cloud after array fusion, a Poisson surface reconstruction method based on point cloud density weighted interpolation is proposed to refine the roadway structure and realize the accurate reconstruction of digital roadway model. The experimental results show that the digital modeling method of millimeter-wave radar array can accurately obtain the information of roadway surrounding rock, the density of roadway point cloud is increased by 22.4%, and the average absolute error percentage of the width and height of the reconstructed roadway model is only 0.82% and 0.72%, which provides a new research method for the reconstruction of underground roadway and an important basis for the digital modeling method of coal mine roadway.

Similar content being viewed by others

characteristics of scientific method of problem solving

A new point cloud processing method unveiled hidden coastal boulders from deep vegetation

characteristics of scientific method of problem solving

Autonomous navigation and collision prediction of port channel based on computer vision and lidar

characteristics of scientific method of problem solving

Step feature line extraction from large-scale point clouds of open-pit mine based on structural characteristics

Introduction.

The development of intelligent coal mine, strongly supported by the state, is crucial for transforming the coal mining industry and enhancing mine safety. Accurate geological information detection is one of the key research and development directions in the construction of smart coal mine. The accurate detection of roadway information and the rapid acquisition of three-dimensional models of roadway are important data sources for geological transparency 1 , 2 . The traditional surveying and mapping technology is mainly based on theodolite, range finder, total station, and GPS, which lacks the details of the three-dimensional model and cannot reflect the complex situation in the coal mine roadway in detail, resulting in the lack of information 3 . The meter-level accuracy of strap-down inertial navigation and ultra-wideband is difficult to meet the positioning needs of the current coal mine 4 . Due to the complex underground environment of coal mine, and the dynamic mining process is accompanied by dynamic changes in space and the generation of a large amount of dust, sensors such as lidar and machine vision cannot adapt to this environment 5 . Coal dust will block the laser and visual sensors, resulting in the failure to scan the coal wall of the working face normally 6 . Ultrasonic and infrared ranging are less affected by the environment, but the disadvantage is that ultrasonic cannot achieve real-time three-dimensional imaging of coal mine roadway. Millimeter-wave radar is more suitable for more complex roadway environment because of its anti-interference and strong penetration. Therefore, it is of great significance to study the intelligent perception and digital modeling of millimeter wave radar in roadway surrounding rock for intelligent mining and accurate positioning of equipment in coal mine.

At present, in terms of point cloud reconstruction technology of three-dimensional roadway, the mainstream 3D roadway construction technology in China is based on laser radar, machine vision and so on. References 7 , 8 , 9 used laser radar to scan the roadway, obtain data information, process, and establish the roadway model. Yang et al. 10 realized the perception of roadway pavement boundary by using visual sensor. The camera and laser sensor can obtain rich texture information and three-dimensional structure information in a good environment. Low illumination and dust in underground environment will affect visual SLAM and laser SLAM reducing the accuracy of mapping 11 . References 12 , 13 used laser SLAM to carry out three-dimensional mapping of coal mine roadway, and the average distance of three-dimensional coordinate drift was large, which could not meet the requirements of actual working conditions of coal mine. Therefore, we consider that millimeter-wave radar has the characteristics of strong penetration and good robustness under complex conditions and unpredictable lighting conditions, so it is suitable for coal mine underground environment perception and map construction. Yonedak et al. 14 modeled the uncertainty of millimeter-wave radar through error propagation to construct maps and locate, and its positioning accuracy in blizzard environment is still centimeter-level. Preprocessing methods such as point cloud filtering and point cloud registration mainly focus on denoising and optimization of the original point cloud data, to achieve more accurate and complete 3D reconstruction data. In the aspect of point cloud filtering processing, the main research is to achieve the ideal noise reduction effect by setting threshold and hybrid filtering algorithm 15 , 16 . However, a single filtering algorithm is no longer applicable, and the noise reduction effect is improved by combining multi-layer filtering. To solve the problems of low detection accuracy of traditional target detection algorithms and slow detection speed of lidar and visual fusion algorithms, Wang et al. 17 proposed a vehicle target detection algorithm based on the fusion of laser lidar and millimeter-wave radar. In the aspect of point cloud registration, the coarse and fine registration of point cloud is mainly realized by extracting overlapping area and transformation algorithm 18 , 19 , 20 . Aiming at the sparse feature problem of millimeter-wave point cloud, the traditional ICP algorithm cannot achieve the ideal accuracy directly, and the registration algorithm needs to be improved by combining the characteristics of millimeter-wave point cloud. Radar point cloud data fusion is the key to solve the problem of radar sparseness. Lin et al. 21 realized the fusion of multi-frame millimeter-wave point clouds through the nearest point iteration method to improve the sparse problem of radar point clouds. For the complex and closed narrow typical coal mine roadway environment, the millimeter-wave radar array can not only overcome the harsh environment, but also solve the problem of sparse point cloud and limited detection range of single millimeter-wave radar. The sparse point cloud of millimeter-wave radar is not suitable for the reconstruction algorithm of dense point cloud (such as laser). It is necessary to improve the reconstruction algorithm which is more suitable for the sparse feature of point cloud. Kazhdan et al. 22 proposed a shielded Poisson surface reconstruction algorithm to interpolate the input point cloud data and used the multi-grid algorithm to solve it, which can keep the sparse point structure of the system unchanged. Xu et al. 23 proposed an adaptive bandwidth Gaussian kernel density estimator, which is beneficial to remove noise and outliers in the process of 3D reconstruction. Gao et al. 24 proposed a Poisson reconstruction algorithm based on improved isosurface extraction, which effectively eliminated the problem of surface holes and misconnected surface features. The research shows that Poisson surface reconstruction is friendly to the integrity of sparse point cloud structure, but the sparse and uneven point cloud will still affect the reconstruction effect and lead to holes and other problems.

In summary, we mainly study the digital modelling method of roadway based on millimeter-wave radar array. This method preprocesses the point cloud data obtained by millimeter-wave radar, such as filtering and registration, to solve the problem of environmental interference and point cloud registration of single millimeter-wave radar. Through the radar array data association fusion algorithm, the limitation of single radar detection is solved, and the accuracy and density of point cloud data of roadway construction are improved. Finally, the point cloud Poisson surface reconstruction algorithm is improved by point cloud density weighted interpolation, to realize the homogenization and densification of point cloud, further improve the accuracy of surface reconstruction, and refine and approximate the real roadway structure. Through the analysis and processing of millimeter-wave radar data, the real-time perception and accurate modelling of coal mine roadway are realized. It provides accurate spatial information and geological characteristics and provides important support for intelligent mining and accurate positioning of equipment in coal mine.

Method and principle

Coal mine tunnel space is narrow, low illumination, high temperature and humidity. There is dust interference, and the dynamic change is large. Aiming at the complex environment of coal mine roadway and the difficulty of accurate perception and real-time accurate modelling, we propose a digital modelling method of coal mine roadway based on millimeter-wave radar array. It includes three parts: preprocessing of millimeter-wave point cloud from coal mine roadway, data fusion of millimeter-wave radar array, and digital modelling and optimization of roadway.As shown in Fig.  1 , which is the technical route of digital modelling scheme of coal mine roadway based on millimeter-wave radar array.

figure 1

Technical route of digital modelling scheme of coal mine roadway based on millimeter wave radar array.

Point cloud preprocessing: Point cloud filtering and registration are commonly used preprocessing methods for point cloud reconstruction. Point cloud data collected by millimeter-wave radar are subjected to multi-layer filtering noise reduction and dynamic sub-graph registration to remove interference points caused by factors such as noise, clutter and multipath. Retain high-precision point cloud data that can accurately express the surrounding rock environment of the roadway and provide accurate data support for roadway reconstruction.

Millimeter wave radar array data fusion: According to the radar position relationship, a unified coordinate system is established through coordinate transformation. The ICP point cloud registration algorithm is improved by using the non-uniformity characteristics of millimeter-wave point cloud. The local roadway data scanned by each radar is fused to obtain complete roadway surrounding rock information.

Digital modelling and optimization of roadway: the weight distribution of point cloud is realized by calculating the density of point cloud blocks, and the Poisson surface reconstruction model is optimized by filling point cloud with linear interpolation. Homogenization and densification processing accurately reflect the detailed characteristics and geometric structure of roadway.

Millimeter-wave point cloud preprocessing of coal mine roadway

The three-dimensional point cloud collected underground by the millimeter-wave radar is sparse and uneven, which is susceptible to noise pollution, so that the roadway point cloud data contains outliers or artifacts. Therefore, it is necessary to preprocess the collected raw data by filtering, noise reduction and registration, to reduce the difficulty of subsequent data fusion and improve the accuracy of 3D point cloud reconstruction on the surface of coal mine roadway. Therefore, the millimeter-wave point cloud preprocessing of coal mine roadway is realized by multi-layer filtering noise reduction processing and dynamic sub-graph registration. Figure  2 is the roadway point cloud preprocessing scheme.

figure 2

Coal mine roadway point cloud preprocessing.

  • Multi-layer filtering noise reduction

Millimeter-wave radar signals are often affected by clutter, noise and multipath. Through multi-layer filtering and noise reduction processing, accurate data are provided for subsequent data registration, fusion calculation, and reconstruction. Considering the physical characteristics of millimeter-wave point cloud, combined with the actual coal mine roadway environment. The point cloud data is accurately processed by multi-layer algorithms such as conditional filtering, straight-through filtering, and local plane fitting filtering. The specific filtering process is shown in Fig.  3 .

figure 3

Multi-layer filtering flow chart.

Conditional filtering

Since the millimeter-wave data contains the information of the target’s distance, coordinates, speed and reflection intensity, millimeter-wave radar has a high sensitivity to the electromagnetic characteristics of different materials. The first step is to carry out conditional filtering according to the strength information. The electromagnetic characteristics of coal, rock, metal, and other materials in coal mine roadway are different to determine the strength threshold. The point cloud of the strength range of coal and rock is retained, and the point cloud data outside the threshold range is eliminated. The retained point cloud data is subjected to straight-through filtering to achieve the elimination of discrete points.

The radar output power is P t , and the signal is transmitted by the antenna of the transmitter. To detect the directivity better, the signal is concentrated and radiated, and the gain G is increased in the directivity of the transmitting antenna. The target object at the distance of R from the radar captures part of the power and reflects back to the receiving antenna. The scattering cross-sectional area of the object is σ , and the effective receiving area is A e . The echo power P r received by the radar is:

Straight-through filtering

The point cloud generated by noise, clutter and other factors is relatively isolated and scattered, and the point cloud data after conditional filtering is straight-through filtering. Considering the positional relationship between the point cloud and the size characteristics of the rectangular roadway shape, the three-dimensional roadway coordinate threshold of the point cloud is determined.

Select any point m i of the point cloud set M , the point cloud m i coordinate is \((x_{i} ,y_{i} ,z_{i} )\) , and the formula is used to determine whether the coordinate of m i is within the threshold range.

where \(x_{\min }\) , \(x_{\max }\) represent the coordinate range of the point cloud in the x direction; \(y_{\min }\) , \(y_{\max }\) represent the coordinate range of the point cloud in the y direction; \(z_{\min }\) , \(z_{\max }\) represent the coordinate range of the point cloud in the z direction. During the acquisition process, the device may have a motion offset. To avoid mistakenly deleting the target point cloud, it is necessary to set the margin parameters \(\Delta x,\Delta y,\Delta z\) in the three-dimensional coordinates. The outliers and interference points in the point cloud set are removed, the point cloud data within the range is retained, and the local plane fitting filtering is performed to achieve the filtering of near-noise points.

Local plane fitting filtering

The straight-through filtering avoids deleting the target point cloud exceeding the roadway threshold due to the movement of the device by setting the margin parameters as a whole but does not consider the near-point noise of the local point cloud. The k nearest neighbor point set of each data point is searched, and the least square method is used to fit the best plane of the point set. The general expression of the local plane equation is:

For the k nearest neighbor point set of the query point, the point coordinates are \((x_{i} ,y_{i} ,z_{i} )\) , where \(i = 1,2, \cdot \cdot \cdot ,k\) . By calculating the distance \(D_{i}\) from the nearest neighbor set to the fitting plane:

The distance from all data points to the corresponding local fitting plane is calculated and the mean value is obtained. The points whose distance from the data points to the local fitting plane is greater than the mean value is regarded as noise points and removed. The selection of the threshold is determined by the average distance from the k nearest neighbor set of all data points to the fitting plane, which improves the adaptive ability of the threshold selection.

Through the above multi-layer filtering, compared with the single filtering algorithm, the outliers, isolated points, and false targets generated by noise, clutter and multipath can be eliminated, to retain the real point cloud with stable reflection characteristics in the environment.

Dynamic sub-graph registration

There is a repeated scanning area between adjacent point cloud frames, which is usually used as a key feature of point cloud registration. Due to the sparseness of the millimeter wave point cloud, the matching characteristics between adjacent frames are not obvious. It is not possible to simply calculate the matching parameters R(θ) and t (x, y) through the nearest neighbor principle to match the current frame with the adjacent frame, and then realize the splicing between the point cloud frames. With the superposition of the number of millimeter wave radar point cloud frames, the contour of the collected roadway surrounding rock gradually appears, and the three-dimensional features are more accurate. Multi-frame point cloud data is selected to form a dynamic sub-graph as the overall calculation. Compared with single-frame point cloud, repeated features are added between sub-graphs, which can not only improve the accuracy of point cloud registration, but also effectively express the feature information of surrounding rock of coal mine roadway. Through the coarse registration and precise registration of the dynamic sub-graph, the registration of the single millimeter wave radar point cloud data of the roadway is realized.

Coarse registration

In the process of data acquisition, there are changes in position or attitude between data frames. It is necessary to register point cloud frames. The timestamp recorded in the data source (data acquisition time) is used for sorting. The displacement L is calculated by data acquisition frequency and translational velocity relation to achieve coarse registration.

With the increase of the moving time T of millimeter-wave radar, the position of the scanning roadway moves forward. The data at different times are spliced to form the point cloud data information of the whole roadway.

Point cloud data in t n time:

where \(\left( {\Delta x,\Delta y,\Delta z} \right)\) is the amount of data change from t n-1 to t n .

By splicing the radar data collected at different times, the formula is as follows:

The complete data roadway information is obtained and stitched as shown in Fig.  4 .

figure 4

Data frame splicing schematic diagram.

Precise registration

Selecting multi-frame point cloud data to merge into a sub-graph after reaching a certain number of point clouds, and using the number of dynamic frames to ensure that the characteristics of the point cloud sub-graph are more obvious. The dynamic sub-graph registration model is established to match the point cloud sub-graph. The point cloud data of the sub-graph is inserted into the optimal position on the previous sub-graph by establishing the least square model. The specific operation is shown in the formula ( 7 ). By continuously matching new data, the multi-frame roadway point cloud data is used to construct a local roadway subgraph, and the subgraph overlap part is used to register the roadway data, to realize the splicing of the roadway point cloud data. The principle is shown in Fig.  5 .

where \(\xi \oplus P_{k}\) is the corresponding point cloud set when the current sub-graph is converted to the previous sub-graph; \(\xi\) is the coordinate of the previous sub-graph; \(P_{k}\) is the point cloud data of the current sub-graph; \(R(\theta )\) is the rotation matrix; \(t(x,y)\) is the translation vector.

figure 5

Schematic diagram of dynamic sub-graph registration principle.

The Euclidean distance sub-graph is established, and the optimal matching is achieved by solving the minimum value of the Euclidean distance sum. The formula is as follows:

where S is the Euclidean distance value, and N is the number of matches between the current sub-graph and the previous sub-graph point cloud.

Millimeter wave radar array data fusion

A single millimeter-wave radar can provide a limited range of roadway environmental information. Multi-millimeter-wave radar information fusion can greatly improve the range of environmental perception. The data fusion of millimeter-wave radar array is a complex problem. It is necessary to solve the matching roadway point cloud data from the detection data obtained by each millimeter-wave radar. Through the technology of coordinate transformation and data fusion, the roadway area scanned by different millimeter-wave radars is fused and spliced, so as to obtain the complete information of roadway surrounding rock.

Millimeter-wave radar array acquisition system

Before the spatial synchronization of multi-millimeter-wave radar, it is necessary to establish the coordinate system of each millimeter-wave radar, and then establish a unified coordinate system according to the spatial distribution relationship. The distribution position of millimeter-wave radars is shown in Fig.  6 . The upper and lower radar height difference L, the upper millimeter-wave radar scans the top of the roadway, and the lower millimeter-wave radar scans the side wall of the roadway.

figure 6

Schematic diagram of radar installation location.

Up-down millimeter-wave radar data fusion

Due to the limitations of single radar scanning, the millimeter-wave radar array is used to scan the roadway, and each millimeter-wave radar is responsible for scanning its own area. Since each point cloud image is composed of multiple point clouds, the point cloud data of different millimeter-wave radars are spliced by using the point cloud data relationship at the junction of different radar scanning areas to realize multi-point cloud data fusion.

The overlapping area of the upper and lower millimeter-wave radar scanning area is small. The simple use of the ICP algorithm will fall into the local minimum value due to the lack of the characteristics of the overlapping area 25 . The millimeter-wave point cloud of the roadway collected by the system has inhomogeneity, and the point density is used for coarse registration according to this feature. The roughly overlapping area is determined by the coordinate relationship, and the overlapping area is divided into blocks and the density of K neighborhood points is calculated. Using the point cloud density as a constraint, the violent cycle method is used to perform a double cycle on the upper and lower point cloud images. Traverse the matching point cloud block pairs, calculate their Euclidean distance, and select the closest point cloud block as the search nearest point as the initial position. Figure  7 is the schematic diagram of rough registration principle.

figure 7

Schematic diagram of rough registration principle.

By calculating the optimal transformation for fine registration, the optimal transformation parameters R and T that make the point cloud map correspond to the point cloud registration are calculated. Suppose that P s and P t represent the source point cloud and the target point cloud respectively, the objective optimization function is to minimize the distance between the corresponding points after transformation, that is:

In the case of a known point correspondence, let \({\overline{\text{p}} }_{\text{s}}\) , \({\overline{\text{p}} }_{\text{t}}\) represent the centroids of the source point cloud and the target point cloud, respectively. Let \(\overset{\lower0.5em\hbox{$\smash{\scriptscriptstyle\frown}$}}{p}_{s}^{i} = p_{s}^{i} - \overline{p}_{s}\) \(\overset{\lower0.5em\hbox{$\smash{\scriptscriptstyle\frown}$}}{p}_{t}^{i} = p_{t}^{i} - \overline{p}_{t}\) and \(H = \sum\limits_{i = 1}^{|Ps|} {\hat{p}_{s}^{i} } \hat{p}_{t}^{{i^{T} }}\) . Let H be decomposed to obtain:

Then the optimal rotation of the ICP problem is:

The optimal translation is:

At each iteration, the current optimal transformation parameters \(R_{k}\) , \(T_{k}\) are obtained, and then the transformation is applied to the current source point cloud. The two steps of ' finding the nearest corresponding point' and 'solving the optimal transformation' are continuously iterated until the variation of \(R_{k}\) , \(T_{k}\) is less than a certain value, and the radar data fusion is completed by fine registration.

Roadway digital modelling and optimization

After the data fusion of the top and side wall of the roadway, there are still problems of uneven and sparse point clouds. The surface reconstruction method based on Poisson equation combines the advantages of global fitting surface reconstruction method and local fitting surface reconstruction method 26 . Globally considering the data of all point sets at one time, the hierarchical structure allowed by Poisson 's method supports local basis functions also has good universality for sparse systems. The surface reconstruction of the fused data directly, the sparse or missing part of the data cannot effectively reflect the characteristics of the surrounding rock of the roadway. Aiming at the above problems, a Poisson surface reconstruction method based on point cloud density weighted interpolation is proposed. Figure  8 is the flow chart of Poisson surface reconstruction based on point cloud weighted interpolation.

figure 8

Flow chart of Poisson surface reconstruction based on point cloud weighted interpolation.

Calculate point cloud density

The density of the whole point cloud image is calculated by using the block calculation method. Taking a point cloud block as an example, the average density of the point cloud block is calculated. When calculating the average density of the point cloud block, it is necessary to traverse the whole point cloud block. Assuming that the number of point clouds in the point cloud block is D, \(d(i,j)\) is used to represent the distance between the query point \(P(x_{i,} y_{i} ,z_{i} )\) and its k nearest neighbor point \(Q(x_{j} ,y_{j} ,z_{j} )\) when the ith point is the query point in the point cloud block. \(S_{i}\) is used to represent the minimum distance when the ith point is the query point in the point cloud block. The formulas are as follows:

Then the average distance of the block point cloud is as follows:

The average density of the point cloud is estimated by calculating the average distance of each point of the point cloud block. The smaller the average distance of the point cloud, the denser the point cloud and the greater the density of point cloud blocks. The larger the average distance of the point cloud, the sparser the point cloud and the smaller the point cloud block density.

Point cloud uniform densification

The sparse point cloud blocks are interpolated to realize the upsampling of the ratio k. The formula is as follows:

where is the minimum average distance of point cloud blocks (the maximum point cloud density), and n is the number of low-density point cloud blocks.

In point cloud reconstruction, three-dimensional coordinate information is very important, especially the discrete point cloud information in sparse point cloud blocks. Their absence has a great influence on the overall structure. To ensure the integrity of the roadway point cloud data, the three-dimensional information of the discrete point cloud is fully utilized, and the density information is converted into distance calculation. The density weighting method is used to distribute the weight of point cloud, improve the weight of discrete points, further restore the surrounding rock information at low density, and improve the accuracy of millimeter-wave point cloud construction of roadway surrounding rock. Figure  9 is the point cloud block density weighted interpolation principle diagram.

figure 9

Block density weighted interpolation principal diagram of point cloud.

In the figure, the point cloud block density of \(p\) point is \(\rho_{p}\) ; the point cloud block density of \(q\) point is \(\rho_{q}\) ; the black point represents the detection point, the red point represents the point to be interpolated; \(d_{q}\) and \(d_{p}\) represents the distance from the detection point to the interpolation point.

The specific steps of the point cloud density weighting method are as follows:

Calculate the point cloud density \(\rho\) of the radius search domain of any point in the point cloud block.

Calculate the distance between interpolation points and data points.

Calculate the density of adjacent point cloud blocks as the weight ratio.

Among them, the weight ratio is inversely proportional to the density, and the sum of the weight ratios is the upsampling ratio.

By using the point cloud linear interpolation module of the density weighting method, the point cloud model can be upsampled to obtain a gridded point cloud model with uniform density and high quality.

Three-dimensional surface reconstruction of uniformly dense point cloud

To refine the structure, reduce the mesh and improve the resolution, it is necessary to increase the number of mesh model patches in the same area. The adaptive octree representation implicit function is used to divide the whole function space, and then the Poisson system is solved. The cubic linear interpolation method is used to allocate the eight adjacent nodes of the current node. The vector field of the surface gradient domain represented by the indicator function is approximately:

where \(Ngbr_{D} (s)\) represents the eight adjacent nodes of the current node \(s.p\) . The depth values of the eight nodes should be \(D\) , and \(\{ \alpha_{o,s} \}\) is used as the weight of the interpolation.

Under the premise of known \(\vec{V}\) , the Poisson equation \(\Delta \chi = \nabla \cdot \vec{V}\) is inversely solved to obtain the indicator function \(\chi\) of the surface. In order to ensure that \(\Delta \chi\) and \(\nabla \cdot \vec{V}\) are in the same function space, the mapping of \(\Delta \chi\) and \(\nabla \cdot \vec{V}\) in the function space is closest. That is:

The process of reconstructing the surface is actually the process of isosurface extraction. To obtain the surface \(\partial M\) corresponding to the input point set, a threshold value of the isosurface needs to be selected first. The isosurface corresponding to the selected equivalence should contain as many sampling points as possible. The position of all points is used as the parameter estimation result to obtain the average value and extract the isosurface.

To verify the effectiveness of the algorithm we proposed, the three-dimensional point cloud data scanned by the millimeter-wave radar array system is used as the reconstruction data in the coal mine experimental roadway. The upper millimeter-wave radar is used to scan the top of the roadway, and the lower millimeter-wave radar is used to scan the side of the roadway, to scan the complete surrounding rock information of the roadway. The width of the laboratory simulation roadway is 3.15 m, the height is 2.6 m, the lower millimeter-wave radar is set to be 1 m high. The upper millimeter-wave radar is set to be 1.7 m high. The main parameters of the millimeter-wave radar are shown in Table 1 .

Multi-layer filtering experiment of roadway point cloud

The original roadway point cloud data is processed by multi-layer noise reduction filtering to eliminate outliers, isolated points and false targets caused by noise, clutter and multipath. The original point cloud data is shown in Fig.  10 a. According to the intensity information of the point cloud, the intensity threshold range is determined, and the conditional filtering processing results are shown in Fig.  10 b. It can be observed that some point clouds which do not meet the requirements are removed. According to the positional relationship between the point clouds and the size characteristics of the rectangular roadway shape, the straight-through filtering processing results are shown in Fig.  10 c. The interference of equipment and personnel inside the roadway is filtered out, and the shape contour of the rectangular roadway is preliminarily displayed. The k nearest neighbor point set of the point cloud is used to fit the local plane, and the distance from the query point to the fitting plane is calculated. The removal of the near noise points is completed by setting a certain threshold. The results of plane fitting filtering are shown in Fig.  10 d. The point cloud of the surrounding rock of the roadway is more concentrated, the range is reduced, and it is easy to register and fit later.

figure 10

Point cloud filtering effect diagram.

Roadway point cloud dynamic sub-graph registration experiment

The amount of point cloud data after filtering and noise reduction is significantly reduced. The millimeter-wave point cloud is sparse and the point cloud registration calculation is small, but there are also problems that the features are not obvious. Using the method of constructing dynamic sub-graph registration, we must first determine the data specification of the sub-graph. According to the experiment, we select the number of point clouds to be 1000 as the minimum data for constructing the dynamic sub-graph. The roadway point clouds of different lengths were selected for dynamic sub-graph registration experiments. Figure  11 is the side wall registration effect diagram and Fig.  12 is the roadway roof registration effect diagram. The point cloud features of the side and top of the roadway after registration are more prominent with the increase of data volume, which is conducive to the overall fusion reconstruction of the roadway in the later stage.

figure 11

Side wall registration effect diagram.

figure 12

Roadway roof registration effect diagram.

Roadway point cloud data fusion experiment

Since each point cloud image is composed of multiple point clouds, the point cloud data of different millimeter-wave radars are spliced by using the point cloud data relationship at the junction of different radar scanning areas to realize multi-point cloud data fusion. Firstly, according to the coordinate transformation relationship, the position of the radar in the same coordinate system is shown in Fig.  13 (left). It is obvious that the point cloud image is not aligned. After data fusion processing, combined with the ICP registration algorithm of point cloud density constraint, the registration effect is as shown in Fig.  13 (right), and the registration effect is obviously improved, which improves the accuracy of the subsequent three-dimensional reconstruction of roadway point cloud.

figure 13

Roadway point cloud data fusion effect diagram.

Roadway point cloud optimization and reconstruction experiment

After the data fusion of roadway roof and side wall, there are still problems of uneven and sparse point cloud. If the surface reconstruction of the fused data is carried out directly, the sparse part cannot effectively reflect the characteristics of the surrounding rock of the roadway. Based on the point cloud density weighted interpolation, the Poisson surface reconstruction effect is optimized, and the average point cloud density is increased by 22.4%, which effectively improves the inhomogeneity and sparsity of the point cloud. The optimized Poisson surface reconstruction perfectly completes the missing roadway. Table 2 shows the comparison of average point cloud density before and after algorithm optimization. The Poisson surface reconstruction effect is optimized based on point cloud density weighted interpolation. Figure  14 is the point cloud density weighted interpolation optimization point cloud diagram, which effectively improves the inhomogeneity and sparsity of the point cloud. Figure  15 is the optimized Poisson surface reconstruction comparison diagram, and the optimized Poisson surface reconstruction perfectly completes the missing roadway. Figure  16 is the actual roadway and reconstruction model.

figure 14

Density weighted interpolation optimization point cloud diagram.

figure 15

Optimized Poisson surface reconstruction comparison diagram.

figure 16

Actual roadway and reconstruction model.

Error analysis

Considering the systematic error of millimeter-wave radar array acquisition, it is relatively reasonable to calculate the width and height information of the roadway as the index of measurement error through data processing. In the scanned simulated roadway, 68 groups of roadway data with fixed length were intercepted, and the average value of width and height in this range was compared with the actual roadway size. Table 3 is the analysis of some roadway errors in the experiment, and the overall roadway error results are shown in Fig.  17 .

figure 17

Overall roadway error results. ( a ) Width error, ( b ) Altitude error.

The mean absolute error (MAE) can effectively reflect the overall error of the roadway. The calculation formula is as follows:

According to the error results, the average absolute error percentage of the width is 0.82%, and the average absolute error percentage of the height is 0.72%. In the case of real-time dynamic measurement, the average error of measurement accuracy is within 3 cm. In complex environment, the accuracy of millimeter-wave radar scanning roadway is obviously better than that of lidar. At the same time, according to the requirements of the coal mine measurement regulations, the allowable deviation of the mine measurement level is 1000 mm and the elevation error is 300 mm. The horizontal error and elevation error of the roadway measurement and reconstruction we proposed meet the requirements. Therefore, the digital modelling method of coal mine roadway based on millimeter-wave radar array we constructed has high reconstruction accuracy, which provides an important basis and method for realizing the digital construction of coal mine.

Aiming at the current situation of low illumination, multi-dust, and lack of high-precision three-dimensional roadway model in coal mine roadway, a digital modelling method of roadway based on millimeter-wave array is proposed, which is based on the problems of sparse, uneven and difficult feature information extraction of millimeter-wave point cloud in roadway. Through the radar array scanning technology, the three-dimensional point cloud data information of the roadway is quickly obtained. The point cloud filtering noise reduction and point cloud registration method are proposed to improve the accuracy of the radar to obtain the roadway data. The data association and fusion method of radar array improves the integrity of roadway data. The Poisson surface reconstruction model of point cloud density weighted interpolation is constructed, and the accurate reconstruction of coal mine roadway is realized.

A noise reduction method of millimeter-wave radar point cloud data based on multi-layer filtering is proposed, which can effectively filter out invalid points, discrete points and false points caused by noise, clutter and multipath, and obtain accurate three-dimensional point cloud information of roadway surrounding rock.

A dynamic sub-graph registration method based on millimeter-wave point cloud is proposed. It solves the problems of small number of single-frame point clouds, inconspicuous features, and difficult matching of millimeter-wave point clouds, and realizes the point cloud data registration of a single radar.

The improved ICP algorithm for data fusion realizes the fusion and splicing of the scanning area between array radars also improves the integrity of the point cloud data of the surrounding rock of the coal mine roadway.

A Poisson surface reconstruction method with point cloud weighted interpolation is proposed, which utilizes the point cloud density weighting method to assign weights to the point cloud, and the average point cloud density is improved by 22.4%, which solves the problem of inhomogeneity of millimeter-wave point cloud data in the process of reconstruction and improves the denseness of the point cloud.

After the experimental roadway verification, the average absolute error percentage of the width is 0.82%, and the average absolute error percentage of the height is 0.72%. The error of the surface reconstruction of the millimeter-wave point cloud is small, which is better than the reconstruction of the laser radar roadway and can ensure the reconstruction accuracy.

Studying the digital modeling method of coal mine roadway based on millimeter-wave radar array, which provides new ideas and methods for coal mine roadway perception and digital modeling, and has important academic research significance. However, the current algorithm cannot realize real-time 3D roadway reconstruction, which has limitations for field operations. In the later stage, further research should be carried out in real-time reconstruction to meet the actual work needs.

Data availability

The datasets generated during and analysed during the current study are not publicly available due to the data being part of ongoing research but are available from the corresponding author on reasonable request.

Wang, H. J., Liu, Z. B., Lei, X. R., Han, B. S. & Lu, Z. Q. Key technologies and engineering practice of 3D laser scanning in coal mine roadway. Coal Geol. Explor. 50 , 109–117. https://doi.org/10.12363/issn.1001-1986.21.10.0589 (2022).

Article   Google Scholar  

Wang, G. F. & Du, Y. B. Development direction of intelligent coal mine and intelligent mining technology. Coal Sci. Technol. 47 , 1–10. https://doi.org/10.13199/j.cnki.cst.2019.01.001 (2019).

Li, M., Jiang, Z., Jiang, L. F. & Sun, Z. M. Research progress on 3D visualization technology for intelligent mine. Coal Sci. Technol. 49 , 153–162. https://doi.org/10.13199/j.cnki.cst.2021.02.019 (2021).

Wang, G. F., Zhao, G. R. & Ren, H. W. Analysis on key technologies of intelligent coal mine and intelligent mining. J. China Coal Soc. 44 , 34–41. https://doi.org/10.13225/j.cnki.jccs.2018.5034 (2019).

Chen, X. Z. et al. Development of millimeter wave radar imaging and SLAM in underground coal mine environment. J. China Coal Soc. 45 , 2182–2192. https://doi.org/10.13225/j.cnki.jccs.zn20.0316 (2020).

Wang, C. F. & Rong, Y. Concept, architecture and key technologies for transparent longwall face. Coal Sci. Technol. 47 , 156–163. https://doi.org/10.13199/j.cnki.cst.2019.07.019 (2019).

Jin, Z., Wang, Z. L. & Zhang, Z. B. Research on measurement technology of surrounding rock deformation of mine development roadway based on 3D laser scanning. Appl. Laser 40 , 1120–1125. https://doi.org/10.14128/j.cnki.al.20204006.1120 (2020).

Zhang, J., Hu, Z. J. & Liu, X. Y. Application of 3D laser scanning equipment in rapid surveying and mapping of mine tunnel engineering. Min. Technol. 20 , 229–232. https://doi.org/10.13828/j.cnki.ckjs.2020.06.063 (2020).

Guo, L. L., Zhou, D. W., Zhang, D. M. & Zhou, B. H. Research on deformation monitoring and supporting of tunnel based on laser point cloud. Saf. Coal Mines 51 , 178–183. https://doi.org/10.13347/j.cnki.mkaq.2020.08.039 (2020).

Yang, J. Y., Zhao, L. Y. & Liang, Y. X. Study on mine patrol robot environment perception based on machine vision. Coal Technol. 42 , 227–229. https://doi.org/10.13301/j.cnki.ct.2023.09.047 (2023).

Chen, Y. Y., Huo, Z. L., Liu, Z. W. & Zhang, Y. H. Development trend and key technology of coal mine transportation robot in China. Coal Sci. Technol. 48 , 233–242. https://doi.org/10.13199/j.cnki.cst.2020.07.024 (2020).

Xue, G. H., Wei, J. B., Li, R. X. & Cheng, J. LeGO-LOAM-SC: An improved simultaneous localization and mapping method fusing LeGO-LOAM and scan context for underground coalmine. Sensors 22 , 520–520. https://doi.org/10.3390/S22020520 (2022).

Article   ADS   PubMed   PubMed Central   Google Scholar  

Liu, F., Wang, H. W. & Liu, Y. Research on 3D spatial mapping of roadways based on multi-sensor fusion. J. China Coal Soc. https://doi.org/10.13225/j.cnki.jccs.2023.0758 (2024).

Yoneda, K., Hashimoto, N., Yanase, R., Aldibaja, M. & Suganuma, N. Vehicle Localization using 76GHz omnidirectional millimeter-wave radar for winter automated driving. In IEEE Intelligent Vehicles Symposium (IV) , pp. 971–977 (2018).

Zhao, J. Y., Ma, J., Chen, S. H., Guo, L. & Xu, C. F. Scene-orientedkd-tree filter algorithm for point cloud. Laser J. 42 , 74–78. https://doi.org/10.14016/j.cnki.jgzz.2021.11.074 (2021).

Li, L. Y. & Zhu, Y. F. Research on denoising algorithm of point cloud data based on hybrid filtering. Jiangxi Sci. 39 , 525–529. https://doi.org/10.13990/j.issn1001-3679.2021.03.026 (2021).

Wang, H., Liu, M. L., Cai, Y. F. & Chen, L. Vehicle target detection algorithm based on fusion of lidar and millimeter wave radar. J. Jiangsu Univ. (Nat. Sci. Ed.) 42 , 389–394. https://doi.org/10.3969/j.issn.1671-7775.2021.04.003 (2021).

Shui, W. & Zhou, M. An approach for model reconstruction based on multi-view scansregistration. In 2010 International Conference on Audio, Language and ImageProcessing , pp. 601–606 (IEEE, 2010).

Li, J., Qian, F. & Chen, X. Point cloud registration algorithm based on overlapping regionextraction. J. Phys. Conf. Ser. 1634 (1), 012012 (2020).

Magnusson, M., Lilienthal, A. & Duckett, T. Scan registration for autonomous mining vehicles using 3D-NDT. J. Field Robot. 24 , 803–827. https://doi.org/10.1002/rob.v24:10 (2007).

Lin, F. T., Yan, P. P., Zhang, H. & Xu, G. Iterative closest point method for point cloud data processing of millimeter wave radar. J. Signal Process. 39 , 288–297. https://doi.org/10.16798/j.issn.1003-0530.2023.02.010 (2023).

Kazhdan, M. & Hoppe, H. Screened poisson surface reconstruction. ACM Trans. Graph. (TOG) 32 , 1–13. https://doi.org/10.1145/2487228.2487237 (2013).

Xu, Z., Xu, C., Hu, J. & Meng, Z. Robust resistance to noise and outliers: Screened Poisson surface reconstruction using adaptive kernel density estimation. Comput. Graph. 97 , 19–27. https://doi.org/10.1016/J.CAG.2021.04.005 (2021).

Gao, F., Zhou, H. & Huang, C. Improved Poisson reconstruction algorithm based on vector field and isosurface. Laser Optoelectron. Prog. 57 , 171–179. https://doi.org/10.3788/LOP57.101016 (2020).

Yi, J. B., Peng, X., Cao, F., Li, J. & Xie, W. J. Research on point cloud registration algorithm based on multi-scale feature fusion. J. Guangxi Normal Univ. (Nat. Sci. Ed.) 42 , 1–12. https://doi.org/10.16088/j.issn.1001-6600.2023082502 (2024).

Wang, Y. L., Liu, F., Shen, J. X. & Liu, X. X. Research on 3D reconstruction of retina based on monocular vision. China Mech. Eng. 27 , 2477–2481. https://doi.org/10.3969/j.issn.1004-132X.2016.18.011 (2016).

Download references

The funding was supported by National Natural Science Foundation of China, 52374161, 52174150, National Key Research and Development Program of China, 2023YFC2907603, National Key Research Development Program Young Scientists Project of China, 2022YFF0605300, Xi'an Science and Technology Plan Project, 22GXFW 0067, Shaanxi Provincial Department of Education serves local special—industrialization cultivation project, 23JC048, Shaanxi Province "Two Chain" Integrated Enterprise (Institute) Joint Project, 2023-LL-QY-03.

Author information

Authors and affiliations.

School of Mechanical Engineering, Xi’an University of Science and Technology, Xi’an, 710054, China

Xusheng Xue, Xingyun Yang, Jianing Yue, Qinghua Mao, Yihan Qin, Hongwei Ma, Jianxin Yang, Huahao Wan, Enqiao Zhang & Junbiao Qiu

Shaanxi Key Laboratory of Mine Electromechanical Equipment Intelligent Detection and Control, Xi’an, 710054, China

Shaanxi Minster Intelligent Technology Co., Ltd, Xi’an, 710054, China

Xiaopeng Li

Xi’an Heavy Equipment Hancheng Coal Mining Machinery Co., Ltd, Hancheng, 715400, China

Rongquan Wang

You can also search for this author in PubMed   Google Scholar

Contributions

X.X. Writing—review & editing X.Y. Writing—Algorithm research Q.M. and H.M. Project support J.Y., Y.Q., J.Y., H.W., E.Z., and J.Q. Experimental validation and data management X.L. and R.W. Industrial experiments.

Corresponding author

Correspondence to Xingyun Yang .

Ethics declarations

Competing interests.

The authors declare no competing interests.

Additional information

Publisher's note.

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

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution-NonCommercial-NoDerivatives 4.0 International License, which permits any non-commercial use, sharing, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if you modified the licensed material. You do not have permission under this licence to share adapted material derived from this article or parts of it. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by-nc-nd/4.0/ .

Reprints and permissions

About this article

Cite this article.

Xue, X., Yang, X., Yue, J. et al. Digital modelling method of coal-mine roadway based on millimeter-wave radar array. Sci Rep 14 , 18585 (2024). https://doi.org/10.1038/s41598-024-69547-5

Download citation

Received : 13 March 2024

Accepted : 06 August 2024

Published : 10 August 2024

DOI : https://doi.org/10.1038/s41598-024-69547-5

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

  • Radar array
  • Sub-graph registration
  • Data fusion
  • Poisson surface reconstruction

By submitting a comment you agree to abide by our Terms and Community Guidelines . If you find something abusive or that does not comply with our terms or guidelines please flag it as inappropriate.

Quick links

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

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

characteristics of scientific method of problem solving

Information

  • Author Services

Initiatives

You are accessing a machine-readable page. In order to be human-readable, please install an RSS reader.

All articles published by MDPI are made immediately available worldwide under an open access license. No special permission is required to reuse all or part of the article published by MDPI, including figures and tables. For articles published under an open access Creative Common CC BY license, any part of the article may be reused without permission provided that the original article is clearly cited. For more information, please refer to https://www.mdpi.com/openaccess .

Feature papers represent the most advanced research with significant potential for high impact in the field. A Feature Paper should be a substantial original Article that involves several techniques or approaches, provides an outlook for future research directions and describes possible research applications.

Feature papers are submitted upon individual invitation or recommendation by the scientific editors and must receive positive feedback from the reviewers.

Editor’s Choice articles are based on recommendations by the scientific editors of MDPI journals from around the world. Editors select a small number of articles recently published in the journal that they believe will be particularly interesting to readers, or important in the respective research area. The aim is to provide a snapshot of some of the most exciting work published in the various research areas of the journal.

Original Submission Date Received: .

  • Active Journals
  • Find a Journal
  • Proceedings Series
  • For Authors
  • For Reviewers
  • For Editors
  • For Librarians
  • For Publishers
  • For Societies
  • For Conference Organizers
  • Open Access Policy
  • Institutional Open Access Program
  • Special Issues Guidelines
  • Editorial Process
  • Research and Publication Ethics
  • Article Processing Charges
  • Testimonials
  • Preprints.org
  • SciProfiles
  • Encyclopedia

electronics-logo

Article Menu

  • Subscribe SciFeed
  • Recommended Articles
  • Google Scholar
  • on Google Scholar
  • Table of Contents

Find support for a specific problem in the support section of our website.

Please let us know what you think of our products and services.

Visit our dedicated information section to learn more about MDPI.

JSmol Viewer

A new entity relationship extraction method for semi-structured patent documents.

characteristics of scientific method of problem solving

1. Introduction

2. related works, 3. methodology, 3.1. patent document ontology modeling method based on hierarchical clustering and association rules, 3.1.1. concept acquisition, 3.1.2. inter-conceptual relationship extraction, 3.2. patent document entity identification method combining statistical learning and deep learning, 3.2.1. rule dictionary, 3.2.2. vector initialization, 3.2.3. hole convolution neural network, 3.2.4. bigru network layer, 3.2.5. crf inference layer, 3.3. patent document entity relationship extraction method integrating attention mechanism, 4. experiments, 4.1. datasets and implementation details, 4.2. comparative and ablation experiments, 4.3. validation, 5. results and analysis, 5.1. experiment results, 5.2. validity results, 6. conclusions, author contributions, data availability statement, conflicts of interest.

  • Pejic-Bach, M.; Pivar, J.; Krstić, Ž. Big data for prediction: Patent analysis—Patenting big data for prediction analysis. In Big Data Governance and Perspectives in Knowledge Management ; IGI Global: Hershey, PA, USA, 2019; pp. 218–240. [ Google Scholar ]
  • Ma, K.; Tian, M.; Tan, Y.; Qiu, Q.; Xie, Z.; Huang, R. Ontology-based BERT model for automated information extraction from geological hazard reports. J. Earth Sci. 2023 , 34 , 1390–1405. [ Google Scholar ] [ CrossRef ]
  • Puccetti, G.; Giordano, V.; Spada, I.; Chiarello, F.; Fantoni, G. Technology identification from patent texts: A novel named entity recognition method. Technol. Forecast. Soc. Chang. 2023 , 186 , 122160. [ Google Scholar ] [ CrossRef ]
  • Yang, G.; Niu, S.; Dai, B.; Zhang, B.; Li, C.; Jiang, Y. Named entity recognition method of blockchain patent text based on deep learning. In Proceedings of the Third International Conference on Electronic Information Engineering, Big Data, and Computer Technology (EIBDCT 2024), Qingdao, China, 21 February 2024; Volume 13181. [ Google Scholar ]
  • Bhattacharya, K.; Chakrabarti, A. A Knowledge Graph and Rule based Reasoning Method for Extracting SAPPhIRE Information from Text. Proc. Des. Soc. 2023 , 3 , 221–230. [ Google Scholar ] [ CrossRef ]
  • Trappey, A.J.C.; Liang, C.-P.; Lin, H.-J. Using machine learning language models to generate innovation knowledge graphs for patent mining. Appl. Sci. 2022 , 12 , 9818. [ Google Scholar ] [ CrossRef ]
  • Yang, Y.; Li, S. Entity Overlapping Relation Extracting Algorithm based on CNN and BERT. IEEE Access 2024 . [ Google Scholar ] [ CrossRef ]
  • Bai, T.; Guan, H.; Wang, S.; Wang, Y.; Huang, L. Traditional Chinese medicine entity relation extraction based on CNN with segment attention. Neural Comput. Appl. 2022 , 34 , 2739–2748. [ Google Scholar ] [ CrossRef ]
  • Shi, M.; Huang, J.; Li, C. Entity relationship extraction based on BLSTM model. In Proceedings of the 2019 IEEE/ACIS 18th International Conference on Computer and Information Science (ICIS), Beijing, China, 17–19 June 2019; IEEE: Piscataway, NJ, USA, 2019. [ Google Scholar ]
  • Wei, M.; Xu, Z.; Hu, J. Entity relationship extraction based on bi-LSTM and attention mechanism. In Proceedings of the 2021 2nd International Conference on Artificial Intelligence and Information Systems, Chongqing, China, 28–30 May 2021. [ Google Scholar ]
  • Liu, Y.; Zuo, Q.; Wang, X.; Zong, T. Entity relationship extraction based on a multi-neural network cooperation model. Appl. Sci. 2023 , 13 , 6812. [ Google Scholar ] [ CrossRef ]
  • Qiao, B.; Zou, Z.; Huang, Y.; Fang, K.; Zhu, X.; Chen, Y. A joint model for entity and relation extraction based on BERT. Neural Comput. Appl. 2021 , 34 , 3471–3481. [ Google Scholar ] [ CrossRef ]
  • Fan, C. The Entity Relationship Extraction Method Using Improved RoBERTa and Multi-Task Learning. Comput. Mater. Contin. 2023 , 77 , 1719–1738. [ Google Scholar ] [ CrossRef ]
  • Lin, Y.; Ji, H.; Huang, F.; Wu, L. A joint neural model for information extraction with global features. In Proceedings of the 58th Annual Meeting of the Association for Computational Linguistics, Online, 5–10 July 2020. [ Google Scholar ]
  • Nasar, Z.; Jaffry, S.W.; Malik, M.K. Named entity recognition and relation extraction: State-of-the-art. ACM Comput. Surv. 2021 , 54 , 1–39. [ Google Scholar ] [ CrossRef ]
  • Miric, M.; Jia, N.; Kenneth, G. Huang. Using supervised machine learning for large-scale classification in management research: The case for identifying artificial intelligence patents. Strategy Manag. J. 2023 , 44 , 491–519. [ Google Scholar ] [ CrossRef ]
  • Lin, H.; Yan, J.; Qu, M.; Ren, X. Learning dual retrieval module for semi-supervised relation extraction. In Proceedings of the World Wide Web Conference, San Francisco, CA, USA, 13–17 May 2019. [ Google Scholar ]
  • Shang, Y.; Huang, H.Y.; Mao, X.L.; Sun, X.; Wei, W. Are noisy sentences useless for distant supervised relation extraction? In Proceedings of the AAAI Conference on Artificial Intelligence, New York, NY, USA, 7–12 February 2020; Volume 34. [ Google Scholar ]
  • Hong, Y.; Li, J.; Feng, J.; Huang, C.; Li, Z.; Qu, J.; Xiao, Y.; Wang, W. Competition or cooperation? exploring unlabeled data via challenging minimax game for semi-supervised relation extraction. In Proceedings of the AAAI Conference on Artificial Intelligence, Washington, DC, USA, 7–14 February 2023; Volume 37. [ Google Scholar ]
  • Kambhatla, N. Combining lexical, syntactic, and semantic features with maximum entropy models for information extraction. In Proceedings of the ACL Interactive Poster and Demonstration Sessions, Barcelona, Spain, 22 July 2004. [ Google Scholar ]
  • Shan, Z.; Liang, F. Extraction of STEM Knowledge Relationship in Physical Education Course Textbooks Based on KNN. In Proceedings of the 2023 IEEE 6th Eurasian Conference on Educational Innovation (ECEI), Singapore, 3–5 February 2023; IEEE: Piscataway, NJ, USA, 2023. [ Google Scholar ]
  • Hou, W.; Hong, L.; Xu, H.; Yin, W. RoRED: Bootstrapping labeling rule discovery for robust relation extraction. Inf. Sci. 2023 , 629 , 62–76. [ Google Scholar ] [ CrossRef ]
  • Li, P.; Mao, K. Knowledge-oriented convolutional neural network for causal relation extraction from natural language texts. Expert Syst. Appl. 2019 , 115 , 512–523. [ Google Scholar ] [ CrossRef ]
  • Zhou, P.; Shi, W.; Tian, J.; Qi, Z.; Li, B.; Hao, H.; Xu, B. Attention-based bidirectional long short-term memory networks for relation classification. In Proceedings of the 54th Annual Meeting of the Association for Computational Linguistics (Volume 2: Short Papers), Berlin, Germany, 7–12 August 2016. [ Google Scholar ]
  • Schlichtkrull, M.; Kipf, T.N.; Bloem, P.; Van Den Berg, R.; Titov, I.; Welling, M. Modeling relational data with graph convolutional networks. In Proceedings of the Semantic Web: 15th International Conference, ESWC 2018, Heraklion, Greece, 3–7 June 2018; proceedings 15. Springer International Publishing: Cham, Switzerland, 2018. [ Google Scholar ]
  • Zhou, H.; Xu, Y.; Yao, W.; Liu, Z.; Lang, C.; Jiang, H. Global context-enhanced graph convolutional networks for document-level relation extraction. In Proceedings of the 28th International Conference on Computational Linguistics, Online, 8–13 December 2020. [ Google Scholar ]
  • Zhen, Y.; Zheng, L.; Chen, P. Constructing knowledge graphs for online collaborative programming. IEEE Access 2021 , 9 , 117969–117980. [ Google Scholar ] [ CrossRef ]
  • Zhao, T.; Yan, Z.; Cao, Y.; Li, Z. Asking effective and diverse questions: A machine reading comprehension based framework for joint entity-relation extraction. In Proceedings of the Twenty-Ninth International Conference on International Joint Conferences on Artificial Intelligence, Yokohama, Japan, 7–15 January 2021. [ Google Scholar ]
  • Oliveira, L.; Claro, D.B.; Souza, M. DptOIE: A Portuguese open information extraction based on dependency analysis. Artif. Intell. Rev. 2023 , 56 , 7015–7046. [ Google Scholar ] [ CrossRef ]
  • Bhatia, P.; Celikkaya, B.; Khalilia, M.; Senthivel, S. Comprehend medical: A named entity recognition and relationship extraction web service. In Proceedings of the 2019 18th IEEE International Conference On Machine Learning And Applications (ICMLA), Boca Raton, FL, USA, 16–19 December 2019; IEEE: Piscataway, NJ, USA, 2019. [ Google Scholar ]
  • Berahmand, K.; Haghani, S.; Rostami, M.; Li, Y. A new attributed graph clustering by using label propagation in complex networks. J. King Saud Univ.-Comput. Inf. Sci. 2022 , 34 , 1869–1883. [ Google Scholar ] [ CrossRef ]
  • Yuan, L.; Cai, Y.; Wang, J.; Li, Q. Joint multimodal entity-relation extraction based on edge-enhanced graph alignment network and word-pair relation tagging. Proc. AAAI Conf. Artif. Intell. 2023 , 37 , 11051–11059. [ Google Scholar ] [ CrossRef ]
  • Kamateri, E.; Stamatis, V.; Diamantaras, K.; Salampasis, M. Automated single-label patent classification using ensemble classifiers. In Proceedings of the 2022 14th International Conference on Machine Learning and Computing, Guangzhou, China, 18–21 February 2022. [ Google Scholar ]
  • Chen, Y.; Yang, W.; Wang, K.; Qin, Y.; Huang, R.; Zheng, Q. A neuralized feature engineering method for entity relation extraction. Neural Netw. 2021 , 141 , 249–260. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Yan, Y.; Okazaki, N.; Matsuo, Y.; Yang, Z.; Ishizuka, M. Unsupervised relation extraction by mining wikipedia texts using information from the web. In Proceedings of the Joint Conference of the 47th Annual Meeting of the ACL and the 4th International Joint Conference on Natural Language Processing of the AFNLP, Singapore, 2–7 August 2009. [ Google Scholar ]

Click here to enlarge figure

DatasetTraining SetTest Set
Dataset 161,5305541
Dataset 215,229729
ModelsDataset 1Dataset 2
Accuracy/%Recall Rate/%F1/%Accuracy/%Recall Rate/%F1/%
CopyRE6256.558.637.736.437.1
GraphRel63.960.061.944.741.142.9
CopyRRL77.868.172.163.359.961.6
ETL-Span85.372.378.084.382.083.1
CasRel88.788.289.593.490.191.8
MPreA90.293.192.1 91.992.8
MPreA (BERT)91.391.291.193.491.392.3
MPreA (ALBERT)91.991.791.593.291.592.4
MPreA (ELECTRA) 93.6
ModelDataset 1Dataset 2
Accuracy/%Recall Rate/%F1/%Accuracy/%Recall Rate/%F1/%
Model-a91.392.492.193.092.291.3
Model-b92.389.189.393.190.690.4
Model-c91.692.592.893.891.993.1
ModelsDataset 3
Accuracy/%Recall Rate/%F1/%
MPreA92.991.293.7
The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

Zhang, L.; Sun, X.; Ma, X.; Hu, K. A New Entity Relationship Extraction Method for Semi-Structured Patent Documents. Electronics 2024 , 13 , 3144. https://doi.org/10.3390/electronics13163144

Zhang L, Sun X, Ma X, Hu K. A New Entity Relationship Extraction Method for Semi-Structured Patent Documents. Electronics . 2024; 13(16):3144. https://doi.org/10.3390/electronics13163144

Zhang, Liyuan, Xiangyu Sun, Xianghua Ma, and Kaitao Hu. 2024. "A New Entity Relationship Extraction Method for Semi-Structured Patent Documents" Electronics 13, no. 16: 3144. https://doi.org/10.3390/electronics13163144

Article Metrics

Article access statistics, further information, mdpi initiatives, follow mdpi.

MDPI

Subscribe to receive issue release notifications and newsletters from MDPI journals

IMAGES

  1. The scientific method is a process for experimentation

    characteristics of scientific method of problem solving

  2. the scientific method problem

    characteristics of scientific method of problem solving

  3. PPT

    characteristics of scientific method of problem solving

  4. What Is Problem-Solving? Steps, Processes, Exercises to do it Right

    characteristics of scientific method of problem solving

  5. Scientific Method: Definition and Examples

    characteristics of scientific method of problem solving

  6. Scientific Method

    characteristics of scientific method of problem solving

COMMENTS

  1. What is the Scientific Method: How does it work and why is it important

    While the scientific method is versatile in form and function, it encompasses a collection of principles that create a logical progression to the process of problem solving: Define a question: Constructing a clear and precise problem statement that identifies the main question or goal of the investigation is the first step. The wording must ...

  2. The scientific method (article)

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

  3. Scientific Method

    The study of scientific method is the attempt to discern the activities by which that success is achieved. ... present science as problem solving and investigate scientific problem solving as a special case of problem-solving in general. ... historical and philosophical studies have shown how "scientific methods with the characteristics as ...

  4. Using the Scientific Method to Solve Problems

    The scientific method can be used to address any situation or problem where a theory can be developed. Although more often associated with natural sciences, it can also be used to develop theories in social sciences (such as psychology, sociology and linguistics), using both. information is information that can be measured, and tends to focus ...

  5. Scientific method

    scientific method, mathematical and experimental technique employed in the sciences. More specifically, it is the technique used in the construction and testing of a scientific hypothesis. The process of observing, asking questions, and seeking answers through tests and experiments is not unique to any one field of science.

  6. What Are The Steps Of The Scientific Method?

    The scientific method is a process that includes several steps: First, an observation or question arises about a phenomenon. Then a hypothesis is formulated to explain the phenomenon, which is used to make predictions about other related occurrences or to predict the results of new observations quantitatively. Finally, these predictions are put to the test through experiments or further ...

  7. Scientific method

    The scientific method is an empirical method for acquiring knowledge that has characterized the development of science since at least the 17th century. The scientific method involves careful observation coupled with rigorous scepticism, because cognitive assumptions can distort the interpretation of the observation.Scientific inquiry includes creating a hypothesis through inductive reasoning ...

  8. Scientific Method: Definition and Examples

    The scientific method is a series of steps followed by scientific investigators to answer specific questions about the natural world. It involves making observations, formulating a hypothesis, and conducting scientific experiments. Scientific inquiry starts with an observation followed by the formulation of a question about what has been ...

  9. The 6 Scientific Method Steps and How to Use Them

    The number of steps varies, but the process begins with an observation, progresses through an experiment, and concludes with analysis and sharing data. One of the most important pieces to the scientific method is skepticism —the goal is to find truth, not to confirm a particular thought. That requires reevaluation and repeated experimentation ...

  10. The Scientific Method Steps, Uses, and Key Terms

    When conducting research, the scientific method steps to follow are: Observe what you want to investigate. Ask a research question and make predictions. Test the hypothesis and collect data. Examine the results and draw conclusions. Report and share the results. This process not only allows scientists to investigate and understand different ...

  11. The Scientific Method: What Is It?

    The scientific method is a step-by-step problem-solving process. These steps include: ... It's a step-by-step problem-solving process that involves: (1) observation, (2) asking questions, (3 ...

  12. Five Characteristics of the Scientific Method

    The scientific method is the system used by scientists to explore data, generate and test hypotheses, develop new theories and confirm or reject earlier results. ... Five Characteristics of the Scientific Method ... Whether you need help solving quadratic equations, inspiration for the upcoming science fair or the latest update on a major storm ...

  13. The scientific method: steps, examples and characteristics

    The different stages of the scientific method are detailed below: 1. Formulation of a problem: The first step in applying the scientific method is to formulate a problem or question to be investigated. The question asked may be triggered by natural curiosity or by the desire to solve a practical problem. 2. Data collection: Once the problem has ...

  14. The scientific method (article)

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

  15. PDF Scientific Method How do Scientists Solve problems

    The following lesson will allow students to name and explain the steps of the scientific method and use the scientific method to solve a problem. Georgia Performance Standards Characteristics of Science: Habits of Mind: SCSh1. Students will evaluate the importance of curiosity, honesty, openness, and skepticism in science a.

  16. PDF Scientific Method: Characteristics, Types and Aims

    Scientific Method: Characteristics, Types and Aims Research is a process by which one acquires dependable and useful information about a phenomenon or a process. It may be broadly defined "as a systematic inquiry towards ... Problem Solving Problem Solving refers the ability to use knowledge, facts, and data to effectively to solve problems ...

  17. Multiple goals, multiple solutions, plenty of second-guessing and

    Unrealistic, outdated ideas that idealize science can set the public up to distrust scientists and the research process. A philosopher of science describes 3 aspects of how science really gets done.

  18. Problem solving

    TRIZ - Problem-solving tools; Scientific method - is an empirical method for acquiring knowledge that has characterized the development of science. ... complex problems have some typical characteristics, which include: [1] complexity (large numbers of items, interrelations, and decisions) enumerability [clarification needed] heterogeneity ...

  19. PDF The scientific method is a systematic method to problem solving. The

    The scientific method is a systematic method to problem solving. The seven steps in the scientific method are: (!)STATING THE PROBLEM. ... Characteristics of Living Things What characteristics do all living things share? Living things are made up of basic units called cells, are based on a universal genetic ...

  20. Identifying a Scientific Problem

    Identifying a Problem. Picture yourself playing with a ball. You throw it up, and you watch it fall back down. You climb up a tree, and you let the ball roll off a branch and then down to the ...

  21. Scientific Method

    Scientific Method Examples. Following is an example of the scientific method: Growing bean plants: What is the purpose: The main purpose of this experiment is to know where the bean plant should be kept inside or outside to check the growth rate and also set the time frame as four weeks. Construction of hypothesis: The hypothesis used is that ...

  22. Scientific Method Characteristics, Steps-B.ed Notes

    Scientific Method characteristics, Steps-B.ed Notes. August 24, 2022 by Ekramul Hoque. The scientific method may otherwise be called the problem-solving method. Also known as the method of science or method of scientist. The scientific method is the procedure that scientists use in the pursuit of science.

  23. Digital modelling method of coal-mine roadway based on ...

    Firstly, aiming at the problem of complex environmental interference, combined with the characteristics of small amount of single frame data of millimeter-wave point cloud, a multi-layer filtering ...

  24. Sustainability

    The effectiveness and scientific rigor of our proposed methods are demonstrated through solving a new energy vehicle selection problem on an online service platform and conducting comparative analysis. ... it is necessary to develop a scientific decision-making method that leverages the advantages of large-scale consumer satisfaction ratings to ...

  25. A New Entity Relationship Extraction Method for Semi-Structured ...

    Aimed at mitigating the limitations of the existing document entity relation extraction methods, especially the complex information interaction between different entities in the document and the poor effect of entity relation classification, according to the semi-structured characteristics of patent document data, a patent document ontology model construction method based on hierarchical ...