Chemical revolution

Last updated
Geoffroy's 1718 Affinity Table: at the head of each column is a chemical species with which all the species below can combine. Some historians have defined this table as being the start of the chemical revolution. Affinity-table.jpg
Geoffroy's 1718 Affinity Table: at the head of each column is a chemical species with which all the species below can combine. Some historians have defined this table as being the start of the chemical revolution.

In the history of chemistry, the chemical revolution, also called the first chemical revolution, was the reformulation of chemistry during the seventeenth and eighteenth centuries, which culminated in the law of conservation of mass and the oxygen theory of combustion.

Contents

During the 19th and 20th century, this transformation was credited to the work of the French chemist Antoine Lavoisier (the "father of modern chemistry"). [2] However, recent work on the history of early modern chemistry considers the chemical revolution to consist of gradual changes in chemical theory and practice that emerged over a period of two centuries. [3] [4] The so-called scientific revolution took place during the sixteenth and seventeenth centuries whereas the chemical revolution took place during the seventeenth and eighteenth centuries. [5]

Primary factors

Several factors led to the first chemical revolution. First, there were the forms of gravimetric analysis that emerged from alchemy and new kinds of instruments that were developed in medical and industrial contexts. In these settings, chemists increasingly challenged hypotheses that had already been presented by the ancient Greeks. For example, chemists began to assert that all structures were composed of more than the four elements of the Greeks or the eight elements of the medieval alchemists. The Irish alchemist, Robert Boyle, laid the foundations for the Chemical Revolution, with his mechanical corpuscular philosophy, which in turn relied heavily on the alchemical corpuscular theory and experimental method dating back to pseudo-Geber. [6] [7]

Earlier works by chemists such as Jan Baptist van Helmont helped to shift the belief in theory that air existed as a single element to that of one in which air existed as a composition of a mixture of distinct kinds of gasses. [8] Van Helmont's data analysis also suggests that he had a general understanding of the law of conservation of mass in the 17th century. [8] Furthermore, work by Jean Rey in the early 17th century with metals like tin and lead and their oxidation in the presence of air and water helped pinpoint the contribution and existence of oxygen in the oxidation process. [9]

Other factors included new experimental techniques and the discovery of 'fixed air' (carbon dioxide) by Joseph Black in the middle of the 18th century. This discovery was particularly important because it empirically proved that 'air' did not consist of only one substance and because it established 'gas' as an important experimental substance. Nearer the end of the 18th century, the experiments by Henry Cavendish and Joseph Priestley further proved that air is not an element and is instead composed of several different gases. Lavoisier also translated the names of chemical substance into a new nomenclatural language more appealing to scientists of the nineteenth century. Such changes took place in an atmosphere in which the industrial revolution increased public interest in learning and practicing chemistry. When describing the task of reinventing chemical nomenclature, Lavoisier attempted to harness the new centrality of chemistry by making the rather hyperbolic claim that: [10]

We must clean house thoroughly, for they have made use of an enigmatical language peculiar to themselves, which in general presents one meaning for the adepts and another meaning for the vulgar, and at the same time contains nothing that is rationally intelligible either for the one or for the other.

Precision instruments

Much of the reasoning behind Antoine Lavoisier being named the "father of modern chemistry" and the start of the chemical revolution lay in his ability to mathematize the field, pushing chemistry to use the experimental methods utilized in other "more exact sciences." [11] Lavoisier changed the field of chemistry by keeping meticulous balance sheets in his research, attempting to show that through the transformation of chemical species the total amount of substance was conserved. Lavoisier used instrumentation for thermometric and barometric measurements in his experiments, and collaborated with Pierre Simon de Laplace in the invention of the calorimeter, an instrument for measuring heat changes in a reaction. [11] In attempting to dismantle phlogiston theory and implement his own theory of combustion, Lavoisier utilized multiple apparatuses. These included a red-hot iron gun barrel which was designed to have water run through it and decompose, and an alteration of the apparatus which implemented a pneumatic trough at one end, a thermometer, and a barometer. The precision of his measurements was a requirement in convincing opposition of his theories about water as a compound, with instrumentation designed by himself implemented in his research.

Despite having precise measurements for his work, Lavoisier faced a large amount of opposition in his research. Proponents of phlogiston theory, such as Keir and Priestley, claimed that demonstration of facts was only applicable for raw phenomena, and that interpretation of these facts did not imply accuracy in theories. They stated that Lavoisier was attempting to impose order on observed phenomena, whereas a secondary source of validity would be required to give definitive proof of the composition of water and non-existence of phlogiston. [11]

Antoine Lavoisier

The latter stages of the revolution was fuelled by the 1789 publication of Lavoisier's Traité Élémentaire de Chimie (Elements of Chemistry). Beginning with this publication and others to follow, Lavoisier synthesised the work of others and coined the term "oxygen". Antoine Lavoisier represented the chemical revolution not only in his publications, but also in the way he practiced chemistry. Lavoisier's work was characterized by his systematic determination of weights and his strong emphasis on precision and accuracy. [12] While it has been postulated that the law of conservation of mass was discovered by Lavoisier, this claim has been refuted by scientist Marcellin Berthelot. [13] Earlier use of the law of conservation of mass has been suggested by Henry Guerlac, noting that scientist Jan Baptist van Helmont had implicitly applied the methodology to his work in the 16th and 17th centuries. Earlier references of the law of conservation of mass and its use were made by Jean Rey in 1630. [13] Although the law of conservation of mass was not explicitly discovered by Lavoisier, his work with a wider array of materials than what most scientists had available at the time allowed his work to greatly expand the boundaries of the principle and its fundamentals. [13]

Lavoisier also contributed to chemistry a method of understanding combustion and respiration and proof of the composition of water by decomposition into its constituent parts. He explained the theory of combustion, and challenged the phlogiston theory with his views on caloric. The Traité incorporates notions of a "new chemistry" and describes the experiments and reasoning that led to his conclusions. Like Newton's Principia , which was the high point of the Scientific Revolution, Lavoisier's Traité can be seen as the culmination of the Chemical Revolution.

Lavoisier's work was not immediately accepted and it took several decades for it gain momentum. [14] This transition was aided by the work of Jöns Jakob Berzelius, who came up with a simplified shorthand to describe chemical compounds based on John Dalton's theory of atomic weights. Many people credit Lavoisier and his overthrow of phlogiston theory as the traditional chemical revolution, with Lavoisier marking the beginning of the revolution and John Dalton marking its culmination.

Méthode de nomenclature chimique

Antoine Lavoisier, in a collaborative effort with Louis Bernard Guyton de Morveau, Claude Louis Berthollet, and Antoine François de Fourcroy, published Méthode de nomenclature chimique in 1787. [15] This work established a terminology for the "new chemistry" which Lavoisier was creating, which focused on a standardized set of terms, establishment of new elements, and experimental work. Méthode established 55 elements which were substances that could not be broken down into simpler composite parts at the time of publishing. [16] By introducing new terminology into the field, Lavoisier encouraged other chemists to adopt his theories and practices in order to use his terms and stay current in chemistry.

Traité élémentaire de chimie

One of Lavoisier's main influences was Étienne Bonnet, abbé de Condillac. Condillac's approach to scientific research, which was the basis of Lavoisier's approach in Traité, was to demonstrate that human beings could create a mental representation of the world using gathered evidence. In Lavoisier's preface to Traité, he states

It is a maxim universally admitted in geometry, and indeed in every branch of knowledge, that, in the progress of investigation, we should proceed from known facts to what is unknown. ... In this manner, from a series of sensations, observations, and analyses, a successive train of ideas arises, so linked together, that an attentive observer may trace back to a certain point the order and connection of the whole sum of human knowledge. [17]

Lavoisier clearly ties his ideas in with those of Condillac, seeking to reform the field of chemistry. His goal in Traité was to associate the field with direct experience and observation, rather than assumption. His work defined a new foundation for the basis of chemical ideas and set a direction for the future course of chemistry. [18]

Humphry Davy

Humphry Davy was an English chemist and a professor of chemistry at the London's Royal Institution in the early 1800s. [19] There he performed experiments that cast doubt upon some of Lavoisier's key ideas such as the acidity of oxygen and the idea of a caloric element. [19] Davy was able to show that acidity was not due to the presence of oxygen using muriatic acid (hydrochloric acid) as proof. [19] He also proved that the compound oxymuriatic acid contained no oxygen and was instead an element, which he named chlorine. [19] Through his use of electric batteries at the Royal Institution Davy first isolated chlorine, followed by the isolation of elemental iodine in 1813. [19] Using the batteries Davy was also able to isolate the elements sodium and potassium. [19] From these experiments Davy concluded that the forces that join chemical elements together must be electrical in nature. [19] Davy also opposed the idea that caloric was an immaterial fluid, arguing instead that heat was a type of motion. [19]

John Dalton

John Dalton was an English chemist who developed the idea of atomic theory of chemical elements. Dalton's atomic theory of chemical elements assumed that each element had unique atoms associated with and specific to that atom. [19] This was in opposition to Lavoisier's definition of elements which was that elements are substances that chemists could not break down further into simpler parts. [19] Dalton's idea also differed from the idea of corpuscular theory of matter, which believed that all atoms were the same, and had been a supported theory since the 17th century. [19] To help support his idea, Dalton worked on defining the relative weights of atoms in chemicals in his work New System of Chemical Philosophy, published in 1808. [19] His text showed calculations to determine the relative atomic weights of Lavoisier's different elements based on experimental data pertaining to the relative amounts of different elements in chemical combinations. [19] Dalton argued that elements would combine in the simplest form possible. [19] Water was known to be a combination of hydrogen and oxygen, thus Dalton believed water to be a binary compound containing one hydrogen and one oxygen. [19]

Dalton was able to accurately compute the relative quantity of gases in atmospheric air. He used the specific gravity of azotic (nitrogen), oxygenous, carbonic acid (carbon dioxide), and hydrogenous gases as well as aqueous vapor determined by Lavoisier and Davy to determine the proportional weights of each as a percent of a whole volume of atmospheric air. [20] Dalton determined that atmospheric air contains 75.55% azotic gas, 23.32% oxygenous gas, 1.03% aqueous vapor, and 0.10% carbonic acid gas. [20]

Jöns Jacob Berzelius

Jöns Jacob Berzelius was a Swedish chemist who studied medicine at the University of Uppsala and was a professor of chemistry in Stockholm. [19] He drew on the ideas of both Davy and Dalton to create an electrochemical view of how elements combined together. Berzelius classified elements into two groups, electronegative and electropositive depending which pole of a galvanic battery they were released from when decomposed. [19] He created a scale of charge with oxygen being the most electronegative element and potassium the most electropositive. [19] This scale signified that some elements had positive and negative charges associated with them and the position of an element on this scale and the element's charge determined how that element combined with others. [19] Berzelius's work on electrochemical atomic theory was published in 1818 as Essai sur la théorie des proportions chimiques et sur l'influence chimique de l'électricité. [19] He also introduced a new chemical nomenclature into chemistry by representing elements with letters and abbreviations, such as O for oxygen and Fe for iron. Combinations of elements were represented as sequences of these symbols and the number of atoms were represented at first by superscripts and then later subscripts. [19]

Related Research Articles

<span class="mw-page-title-main">Antoine Lavoisier</span> French nobleman and chemist (1743–1794)

Antoine-Laurent de Lavoisier, also Antoine Lavoisier after the French Revolution, was a French nobleman and chemist who was central to the 18th-century chemical revolution and who had a large influence on both the history of chemistry and the history of biology.

<span class="mw-page-title-main">History of atomic theory</span> History of atomic physics

Atomic theory is the scientific theory that matter is composed of particles called atoms. The definition of the word "atom" has changed over the years in response to scientific discoveries. Initially, it referred to a hypothetical concept of there being some fundamental particle of matter, too small to be seen by the naked eye, that could not be divided. Then the definition was refined to being the basic particles of the chemical elements when chemists observed that the elements seemed to combine with each other by a basic quantity. Then physicists discovered that these particles had an internal structure of their own and therefore perhaps did not deserve to be called "atoms", but renaming atoms would have been impractical by that point.

Chemistry is the scientific study of the properties and behavior of matter. It is a physical science within the natural sciences that studies the chemical elements that make up matter and compounds made of atoms, molecules and ions: their composition, structure, properties, behavior and the changes they undergo during reactions with other substances. Chemistry also addresses the nature of chemical bonds in chemical compounds.

<span class="mw-page-title-main">Claude Louis Berthollet</span> French chemist (1748–1822)

Claude Louis Berthollet was a Savoyard-French chemist who became vice president of the French Senate in 1804. He is known for his scientific contributions to theory of chemical equilibria via the mechanism of reverse chemical reactions, and for his contribution to modern chemical nomenclature. On a practical basis, Berthollet was the first to demonstrate the bleaching action of chlorine gas, and was first to develop a solution of sodium hypochlorite as a modern bleaching agent.

<span class="mw-page-title-main">Phlogiston theory</span> Superseded theory of combustion

The phlogiston theory, a superseded scientific theory, postulated the existence of a fire-like element dubbed phlogiston contained within combustible bodies and released during combustion. The name comes from the Ancient Greek φλογιστόνphlogistón, from φλόξphlóx (flame). The idea of a phlogistic substance was first proposed in 1667 by Johann Joachim Becher and later put together more formally in 1703 by Georg Ernst Stahl. Phlogiston theory attempted to explain chemical processes such as combustion and rusting, now collectively known as oxidation. The theory was challenged by the concomitant weight increase and was abandoned before the end of the 18th century following experiments by Antoine Lavoisier in the 1770s and by other scientists. Phlogiston theory led to experiments that ultimately resulted in the identification and naming (1777) of oxygen by Joseph Priestley.

<span class="mw-page-title-main">Carl Wilhelm Scheele</span> Swedish German chemist who discovered oxygen (1742–1786)

Carl Wilhelm Scheele was a Swedish German pharmaceutical chemist.

<span class="mw-page-title-main">Chemical symbol</span> Abbreviations used in chemistry

Chemical symbols are the abbreviations used in chemistry, mainly for chemical elements; but also for functional groups, chemical compounds, and other entities. Element symbols for chemical elements normally consist of one or two letters from the Latin alphabet and are written with the first letter capitalised.

<span class="mw-page-title-main">Antoine-François de Fourcroy</span> French chemist

Antoine François Fourcroy was a French chemist and a contemporary of Antoine Lavoisier. Fourcroy collaborated with Lavoisier, Guyton de Morveau, and Claude Berthollet on the Méthode de nomenclature chimique, a work that helped standardize chemical nomenclature.

The caloric theory is an obsolete scientific theory that heat consists of a self-repellent fluid called caloric that flows from hotter bodies to colder bodies. Caloric was also thought of as a weightless gas that could pass in and out of pores in solids and liquids. The "caloric theory" was superseded by the mid-19th century in favor of the mechanical theory of heat, but nevertheless persisted in some scientific literature—particularly in more popular treatments—until the end of the 19th century.

<span class="mw-page-title-main">Amount of substance</span> Extensive physical property

In chemistry, the amount of substance (symbol n) in a given sample of matter is defined as a ratio (n = N/NA) between the number of elementary entities (N) and the Avogadro constant (NA). The entities are usually molecules, atoms, or ions of a specified kind. The particular substance sampled may be specified using a subscript, e.g., the amount of sodium chloride (NaCl) would be denoted as nNaCl. The unit of amount of substance in the International System of Units is the mole (symbol: mol), a base unit. Since 2019, the value of the Avogadro constant NA is defined to be exactly 6.02214076×1023 mol−1. Sometimes, the amount of substance is referred to as the chemical amount or, informally, as the "number of moles" in a given sample of matter.

<span class="mw-page-title-main">History of chemistry</span>

The history of chemistry represents a time span from ancient history to the present. By 1000 BC, civilizations used technologies that would eventually form the basis of the various branches of chemistry. Examples include the discovery of fire, extracting metals from ores, making pottery and glazes, fermenting beer and wine, extracting chemicals from plants for medicine and perfume, rendering fat into soap, making glass, and making alloys like bronze.

Chemical nomenclature is a set of rules to generate systematic names for chemical compounds. The nomenclature used most frequently worldwide is the one created and developed by the International Union of Pure and Applied Chemistry (IUPAC).

<span class="mw-page-title-main">Marie-Anne Paulze Lavoisier</span> French chemist and artist

Marie-Anne Pierrette Paulze Lavoisier, later Countess von Rumford, was a French chemist and noblewoman. Madame Lavoisier's first husband was the chemist and nobleman Antoine Lavoisier. She acted as his laboratory companion, using her linguistic skills to write up his work and bring it to an international audience. She also played a pivotal role in the translation of several scientific works, and was instrumental to the standardization of the scientific method.

<span class="mw-page-title-main">Timeline of chemistry</span> List of events in the history of chemistry

This timeline of chemistry lists important works, discoveries, ideas, inventions, and experiments that significantly changed humanity's understanding of the modern science known as chemistry, defined as the scientific study of the composition of matter and of its interactions.

<span class="mw-page-title-main">Pneumatic chemistry</span> Very first studies of the role of gases in the air in combustion reactions

In the history of science, pneumatic chemistry is an area of scientific research of the seventeenth, eighteenth, and early nineteenth centuries. Important goals of this work were the understanding of the physical properties of gases and how they relate to chemical reactions and, ultimately, the composition of matter. The rise of phlogiston theory, and its replacement by a new theory after the discovery of oxygen as a gaseous component of the Earth atmosphere and a chemical reagent participating in the combustion reactions, were addressed in the era of pneumatic chemistry.

<span class="mw-page-title-main">Jöns Jacob Berzelius</span> Swedish chemist (1779–1848)

Baron Jöns Jacob Berzelius (Swedish:[jœnsˈjɑ̌ːkɔbbæˈʂěːlɪɵs] was a Swedish chemist. In general, he is considered the last person to know the whole field of chemistry. Berzelius is considered, along with Robert Boyle, John Dalton, and Antoine Lavoisier, to be one of the founders of modern chemistry. Berzelius became a member of the Royal Swedish Academy of Sciences in 1808 and served from 1818 as its principal functionary. He is known in Sweden as the "Father of Swedish Chemistry". During his lifetime he did not customarily use his first given name, and was universally known simply as Jacob Berzelius.

Chemistry: A Volatile History is a 2010 BBC documentary on the history of chemistry presented by Jim Al-Khalili. It was nominated for the 2010 British Academy Television Awards in the category Specialist Factual.

The origin and usage of the term metalloid is convoluted. Its origin lies in attempts, dating from antiquity, to describe metals and to distinguish between typical and less typical forms. It was first applied to metals that floated on water, and then more popularly to nonmetals. Only recently, since the mid-20th century, has it been widely used to refer to elements with intermediate or borderline properties between metals and nonmetals.

<span class="mw-page-title-main">Claudine Picardet</span> French chemist, mineralogist, meteorologist and translator

Claudine Picardet was a chemist, mineralogist, meteorologist and scientific translator. Among the French chemists of the late eighteenth century she stands out for her extensive translations of scientific literature from Swedish, English, German and Italian to French. She translated three books and thousands of pages of scientific papers, which were published as well as circulated in manuscript form. She hosted renowned scientific and literary salons in Dijon and Paris, and was an active participant in the collection of meteorological data. She helped to establish Dijon and Paris as scientific centers, substantially contributing to the spread of scientific knowledge during a critical period in the chemical revolution.

References

  1. Kim, Mi Gyung (2003). Affinity, That Elusive Dream: A Genealogy of the Chemical Revolution. MIT Press. ISBN   978-0-262-11273-4.
  2. The First Chemical Revolution Archived April 26, 2009, at the Wayback Machine – the Instrument Project, The College of Wooster
  3. Eddy, Matthew Daniel; Mauskopf, Seymour H.; Newman, William R. (January 2014). "An Introduction to Chemical Knowledge in the Early Modern World". Osiris. 29 (1): 1–15. doi:10.1086/678110. PMID   26103744.
  4. Florin George Calian. Alkimia Operativa and Alkimia Speculativa. Some Modern Controversies on the Historiography of Alchemy.
  5. Eddy, Matthew D.; Mauskopf, Seymour H.; Newman, William R. (2015). Osiris, Volume 29: Chemical Knowledge in the Early Modern World. University of Chicago Press Journals. ISBN   978-0-226-15839-6.[ page needed ]
  6. Ursula Klein (July 2007). "Styles of Experimentation and Alchemical Matter Theory in the Scientific Revolution". Metascience. 16 (2). Springer: 247–256 [247]. doi:10.1007/s11016-007-9095-8. ISSN   1467-9981. S2CID   170194372.
  7. Niermeier-Dohoney, Justin Robert (2018). A Vital Matter: Alchemy, Cornucopianism, and Agricultural Improvement in Seventeenth-Century England (Thesis). OCLC   1090369187. ProQuest   2161867295.[ page needed ]
  8. 1 2 Ducheyne, Steffen (2008). "A Preliminary Study of the Appropriation of Van Helmont's oeuvre in Britain in Chymistry, Medicine and Natural Philosophy". Ambix. 55 (2): 122–135. doi:10.1179/174582308X255479. ISSN   0002-6980. PMID   19048972. S2CID   38195230.
  9. De Milt, Clara (1953). "The essays of Jean Rey". Journal of Chemical Education. 30 (7): 377. Bibcode:1953JChEd..30..377D. doi: 10.1021/ed030p377.3 . ISSN   0021-9584.
  10. Jaffe, B. (1976). Crucibles: The Story of Chemistry from Alchemy to Nuclear Fission (4th ed.). New York: Dover Publications. ISBN   978-0-486-23342-0.
  11. 1 2 3 Golinski, Jan (1994). "Precision instruments and the demonstrative order of proof in Lavoisier's chemistry". Osiris. 9: 30–47. doi:10.1086/368728. S2CID   95978870.
  12. Levere, Trevor (2001). Transforming Matter. Baltimore, Maryland: The Johns Hopkins University Press. ISBN   0-8018-6610-3.
  13. 1 2 3 Blumenthal, Geoffrey (February 2013). "On Lavoisier's Achievement in Chemistry: On Lavoisier's achievement in chemistry" (PDF). Centaurus. 55 (1): 20–47. doi:10.1111/1600-0498.12001. hdl:1983/205ebdf7-ee96-42db-8687-a1b9eb6575c5.
  14. Eddy, Matthew (2008). The Language of Mineralogy: John Walker, Chemistry and the Edinburgh Medical School, 1750-1800. Ashgate. ISBN   978-0-7546-6332-4.[ page needed ]
  15. Duveen, Denis; Klickstein, Herbert (Sep 1954). "The Introduction of Lavoisier's Chemical Nomenclature into America". Isis. 45 (3): 278–292. doi:10.1086/348339. PMID   13232806. S2CID   32064170.
  16. Guyton de Morveau, Louis-Bernard; Lavoisier, Antoine Laurent; Berthollet, Claude-Louis; Fourcroy, Antoine-François de, comte; Hassenfratz, Jean-Henri; Adet, Pierre-Auguste (1787). Méthode de nomenclature chimique. Paris, France: Chez Cuchet. Archived from the original on 19 April 2019. Retrieved 19 April 2019.{{cite book}}: CS1 maint: multiple names: authors list (link)
  17. Antoine-Laurent Lavoisier, Elements of Chemistry, trans. Robert Kerr (Edinburgh, 1790; facs. reprint New York: Dover, 1965), pp. xv–xvi.
  18. Dear, Peter (2006). The Intelligibility of Nature. The University of Chicago Press. pp. 74–75.
  19. 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 J., Bowler, Peter; Morus, Iwan Rhys (2005). Making modern science: a historical survey. Chicago: University of Chicago Press. ISBN   0226068609. OCLC   56333962.{{cite book}}: CS1 maint: multiple names: authors list (link)
  20. 1 2 Society, Manchester Literary and Philosophical (1805). Memoirs and Proceedings of the Manchester Literary & Philosophical Society: (Manchester Memoirs.).

Further reading