Jump to content

History of atomic theory: Difference between revisions

From Wikipedia, the free encyclopedia
Content deleted Content added
No edit summary
No edit summary
Line 5: Line 5:
Atomic theory began thousands of years ago as a philosophical concept, and in the 19th century achieved widespread scientific acceptance thanks to discoveries in the field of [[stoichiometry]]. The chemists of the era believed the basic units of the elements were also the fundamental particles of nature and named them atoms (derived from the Greek word ''atomos'', meaning "indivisible"). However, around the turn of the 20<sup>th</sup> century, through various experiments with [[electromagnetism]] and [[radioactivity]], physicists discovered that the so-called "indivisible atom" was actually a conglomerate of various subatomic particles (chiefly, [[electrons]], [[protons]] and [[neutrons]]) which can exist separately of each other. In fact, in certain extreme environments such as [[neutron stars]], extreme temperature and pressure prevents atoms from existing at all. The field of science which studies subatomic particles is [[particle physics]], and it is in this field that physicists hope to discover the true fundamental nature of matter.
Atomic theory began thousands of years ago as a philosophical concept, and in the 19th century achieved widespread scientific acceptance thanks to discoveries in the field of [[stoichiometry]]. The chemists of the era believed the basic units of the elements were also the fundamental particles of nature and named them atoms (derived from the Greek word ''atomos'', meaning "indivisible"). However, around the turn of the 20<sup>th</sup> century, through various experiments with [[electromagnetism]] and [[radioactivity]], physicists discovered that the so-called "indivisible atom" was actually a conglomerate of various subatomic particles (chiefly, [[electrons]], [[protons]] and [[neutrons]]) which can exist separately of each other. In fact, in certain extreme environments such as [[neutron stars]], extreme temperature and pressure prevents atoms from existing at all. The field of science which studies subatomic particles is [[particle physics]], and it is in this field that physicists hope to discover the true fundamental nature of matter.



Sarmad butt suk man willys
==Philosophical atomism==
==Philosophical atomism==
{{Main|Atomism}}
{{Main|Atomism}}

Revision as of 11:23, 18 September 2007

This article focuses on historical models of the atom. For a history of the study of how atoms combine to form molecules, see History of the molecule.

In chemistry and physics, atomic theory is a theory of the nature of matter, which states that matter is composed of discrete units called atoms, as opposed to obsolete beliefs that matter could be divided into any arbitrarily small quantity.

Atomic theory began thousands of years ago as a philosophical concept, and in the 19th century achieved widespread scientific acceptance thanks to discoveries in the field of stoichiometry. The chemists of the era believed the basic units of the elements were also the fundamental particles of nature and named them atoms (derived from the Greek word atomos, meaning "indivisible"). However, around the turn of the 20th century, through various experiments with electromagnetism and radioactivity, physicists discovered that the so-called "indivisible atom" was actually a conglomerate of various subatomic particles (chiefly, electrons, protons and neutrons) which can exist separately of each other. In fact, in certain extreme environments such as neutron stars, extreme temperature and pressure prevents atoms from existing at all. The field of science which studies subatomic particles is particle physics, and it is in this field that physicists hope to discover the true fundamental nature of matter.


Philosophical atomism

Indian

Up until the beginning of the 19th century, atomic theory was mainly philosophical and not founded in scientific experimentation. The earliest known theories were developed in ancient India in the 6th century BCE by Kanada, a Hindu philosopher. In Hindu philosophy, the Nyaya and Vaisheshika schools developed elaborate theories on how atoms combined into more complex objects (first in pairs, then trios of pairs), but believed the interactions were ultimately driven by the will of God (specifically, the Hindu Ishvara), and that the atoms themselves were otherwise inactive, without physical properties of their own. By contrast, Jainic philosophy linked the behavior of matter to the nature of the atoms themselves. Each atom, according to Jaina philosophy, has one kind of taste, one smell, one color, and two kinds of touch, though it is unclear what was meant by “kind of touch”. Atoms can exist in one of two states: subtle, in which case they can fit in infinitesimally small spaces, and gross, in which case they have extension and occupy a finite space. Although atoms are made of the same basic substance, they can combine based on their eternal properties to produce any of six “aggregates,” which seem to correspond with the Greek concept of “elements”: earth, water, shadow, sense objects, karmic matter, and unfit matter.[1]

Greek

Leucippus and Democritus, Greek philosophers in the 5th century BCE, presented their own theory of atoms. The Greeks believed that atoms were all made of the same material but had different shapes and sizes, which determined the physical properties of the material. For instance, the atoms of a liquid were thought to be smooth, allowing them to slide over each other.[2] In this line of thought, graphite and diamond would be composed of two different kinds of atoms, but today we know that they're both allotropic forms of carbon.

Islamic

During the 11th century (in the Islamic Golden Age), Islamic atomists developed atomic theories that represent a synthesis of both Greek and Indian atomism. Older Greek and Indian ideas were further developed by Islamic atomists, along with new Islamic ideas, such as the possibility of there being particles smaller than an atom. The most successful form of Islamic atomism was in the Asharite school of philosophy, most notably in the work of the philosopher al-Ghazali (1058-1111). In Asharite atomism, atoms are the only perpetual, material things in existence, and all else in the world is “accidental” meaning something that lasts for only an instant. Nothing accidental can be the cause of anything else, except perception, as it exists for a moment. Contingent events are not subject to natural physical causes, but are the direct result of God’s constant intervention, without which nothing could happen. Thus nature is completely dependent on God, which meshes with other Asharite Islamic ideas on causation, or the lack thereof.[3]

Modern atomic theory

Birth

In the early years of the 19th century, John Dalton developed his atomic theory in which he proposed that each chemical element is composed of atoms of a single, unique type, and that though they are both immutable and indestructible, they can combine to form more complex structures (chemical compounds). How precisely Dalton arrived at his theory is not entirely clear, but nonetheless it allowed him to explain a number of chemistry puzzles he and his contemporaries were pondering at the time.

The first was the law of conservation of mass, formulated by Antoine Lavoisier in 1789, which states that the total mass in a chemical reaction remains constant (that is, the reactants have the same mass as the products).[4] This law suggested to Dalton that matter is fundamentally indestructible.

The second was the law of definite proportions. First proven by the French chemist Joseph Louis Proust in 1799,[5] this law states that if a compound is broken down into its constituent elements, then the masses of the constituents will always have the same proportions, regardless of the quantity or source of the original substance. Proust had synthesized copper carbonate through numerous methods and found that in each case the ingredients combined in the same proportions as they were produced when he broke down natural copper carbonate.

Dalton studied and expanded upon Proust's work to develop the law of multiple proportions: if two elements form more than one compound between them, then the ratios of the masses of the second element which combine with a fixed mass of the first element will be ratios of small integers. One pair of reactions Dalton is believed to have studied involved nitric oxide (NO) and oxygen (O2). In one combination, these gases formed dinitrogen trioxide (N2O3), but when he repeated the combination with twice the amount of oxygen (a ratio of 1:2 - small integers), they instead formed nitrogen dioxide (NO2).

4NO + O2 → 2N2O3

4NO + 2O2 → 4NO2

Dalton also believed atomic theory could explain why water absorbed different gases in different proportions: for example, he found that water absorbed carbon dioxide far better than it absorbed nitrogen. Dalton hypothesized this was due to the differences in mass and complexity of the gases' respective particles. Indeed, carbon dioxide molecules (CO2) are heavier and larger than nitrogen molecules (N2).

Various atoms and molecules as depicted in John Dalton's A New System of Chemical Philosophy (1808).

In 1803 Dalton published his first list of relative atomic weights for a number of substances (though he did not publicly discuss how he obtained these figures until 1808).[6] Dalton estimated the atomic weights according to the mass ratios in which they combined, with hydrogen being the basic unit. However, Dalton did not conceive that with some elements atoms exist in molecules – e.g. pure oxygen exists as O2. He also mistakenly believed that the simplest compound between any two elements is always one atom of each (so he thought water was HO, not H2O)[7]. This, in addition to the crudity of his equipment, resulted in his table being highly flawed. For instance, he believed oxygen atoms were 5.5 times heavier than hydrogen atoms, because in water he measured 5.5 grams of oxygen for every 1 gram of hydrogen and believed the formula for water was HO (an oxygen atom is actually 16 times heavier than a hydrogen atom).

The flaw in Dalton's theory was corrected in 1811 by Amedeo Avogadro. Avogadro had proposed that equal volumes of any two gases, at equal temperature and pressure, contain equal numbers of molecules (in other words, the mass of a gas's particles does not affect its volume).[8] Avogadro's law allowed him to deduce the diatomic nature of numerous gases by studying the volumes at which they reacted. For instance: since two liters of hydrogen will react with just one liter of oxygen to produce two liters of water vapor (at constant pressure and temperature), it meant a single oxygen molecule splits in two in order to form two particles of water. Thus, Avogadro was able to offer more accurate estimates of the atomic mass of oxygen and various other elements, and firmly established the distinction between molecules and atoms.

In 1827, the British botanist Robert Brown observed that dust particles floating in water constantly jiggled about for no apparent reason. In 1905, Albert Einstein theorised that this Brownian motion was caused by the water molecules continuously knocking the grains about, and developed a hypothetical mathematical model to describe it.[9] This model was validated experimentally in 1911 by French physicist Jean Perrin, thus providing additional validation for particle theory (and by extension atomic theory).

Discovery of subatomic particles

Thomson's cathode ray tube in which he observed the deflection of cathode rays by an electric field.

Atoms were thought to be the smallest possible division of matter until 1897 when J.J. Thomson discovered the electron through his work on cathode ray tubes.[10] A cathode ray tube of the kind used by Thomson is a sealed glass container in which two electrodes are separated by a vacuum. When a voltage is applied across the electrodes, cathode rays are generated, creating a glowing patch where they strike the glass at the opposite end of the tube. Through experimentation, Thomson discovered that the rays could be deflected by an electric field (in addition to magnetic fields, which was already known). He concluded that these rays, rather than being waves, were composed of negatively charged particles he called "corpuscles" (they would later be renamed electrons by other scientists).

Thomson believed that the corpuscles emerged from the very atoms of the electrode. He thus concluded that atoms were divisible, and that the corpuscles were their building blocks. To explain the overall neutral charge of the atom, he proposed that the corpuscles were distributed in ring-like structures in a uniform sea or cloud of positive charge; this was the plum pudding model.[11]

Since atoms were found to be actually divisible, physicists later invented the term "elementary particles" to describe indivisible particles.

Discovery of the nucleus

The gold foil experiment
Top: Expected results: alpha particles passing through the plum pudding model of the atom with negligible deflection.
Bottom: Observed results: a small portion of the particles were deflected, indicating a small, concentrated positive charge.

Thomson's plum pudding model was disproved in 1909 by one of his students, Ernest Rutherford, who discovered that most of the mass and positive charge of an atom is concentrated in a very small fraction of its volume, which he assumed to be at the very center.

In the gold foil experiment, Hans Geiger and Ernest Marsden shot alpha particles through a sheet of gold (striking a fluorescent screen that surrounded the foil).[12] Given the very small mass of the electrons, the high momentum of the alpha particles and the unconcentrated distribution of positive charge of the plum pudding model, the experimenters expected all the alpha particles to either pass through without significant deflection or be absorbed. To their astonishment, a small fraction of the alpha particles experienced heavy deflection. This led Rutherford to propose the planetary model of the atom in which pointlike electrons orbited in the space around a massive, compact nucleus—like planets orbiting the Sun.[13]

Discovery of isotopes

In 1913, J.J. Thomson channeled a stream of neon ions through magnetic and electric fields, striking a photographic plate on the other side. He observed two glowing patches on the plate, which suggested two different deflection trajectories. Thomson concluded this was because some of the neon ions had a different mass; thus did he discover the existence of isotopes.[14]

Discovery of nuclear particles

In 1918, Rutherford managed to split the nucleus of the atom when he bombarded nitrogen gas with alpha particles and observed hydrogen nuclei being emitted from the gas. Rutherford concluded that the hydrogen nuclei emerged from the nuclei of the nitrogen atoms themselves.[15] He later found that the positive charge of any atom could always be equated to that of an integer number of hydrogen nuclei. This, coupled with the facts that hydrogen was the lightest element known and that the atomic mass of every other element was roughly equivalent to an integer multiple of hydrogen's atomic mass, led him to conclude hydrogen nuclei were singular particles and a basic constituent of all atomic nuclei: the proton. Further experimentation by Rutherford found that the nuclear mass of most atoms exceeded that of the protons it possessed; he speculated that this surplus mass was composed of hitherto unknown neutrally charged particles.

In 1928, Walter Bothe observed that beryllium emitted an electrically neutral radiation when bombarded with alpha particles. In 1932, James Chadwick exposed various elements to this radiation, and by comparing the energies of recoiling charged particles from the different targets, he deduced that the radiation was composed of electrically neutral particles with a mass similar to that of a proton.[16] Chadwick called these particles "neutrons".

Quantum physical models of the atom

The planetary model of the atom had shortcomings. Firstly, according to the Larmor formula in classical electromagnetism, an accelerating electric charge emits electromagnetic waves; an orbiting charge would steadily lose energy and spiral towards the nucleus, colliding with it in a small fraction of a second. Another phenomenon the model did not explain was why excited atoms only emit light within certain discrete spectra.

The Bohr model of the atom

Quantum theory revolutionized physics at the beginning of the 20th century, when Max Planck and Albert Einstein postulated that light energy is emitted or absorbed in fixed amounts known as quanta (singular, quantum). In 1913, Niels Bohr incorporated this idea into his Bohr model of the atom, in which the electrons could only orbit the nucleus in particular circular orbits with fixed angular momentum and energy, their distances from the nucleus being proportional to their respective energies.[17] Under this model electrons could not spiral into the nucleus because they could not lose energy in a continuous manner; instead, they could only make instantaneous "quantum leaps" between the fixed energy levels.[17] When this occurred, light was emitted or absorbed at a frequency proportional to the change in energy (hence the absorption and emission of light in discrete spectra).[17]

Bohr's model was only able to predict the spectral lines of hydrogen; it couldn't predict those of multielectron atoms. Worse still, as spectrographic technology improved, additional spectral lines in hydrogen were observed which Bohr's model couldn't explain. In 1916, Arnold Sommerfeld added eliptical orbits to the Bohr model to explain the extra emmission lines, but this made the model very difficult to use, and it still couldn't explain complex atoms.

In 1924, Louis de Broglie proposed that all objects—particularly subatomic particles such as electrons — exhibit a degree of wave-like behavior. Erwin Schrodinger, fascinated by this idea, explored whether or not the movement of an electron in an atom could be better explained as a wave rather than as a particle. Schrodinger's equation, published in 1926,[18] describes an electron as a wavefunction instead of as a point particle, and it elegantly predicted many of the spectral phenomena Bohr's model failed to explain. Although this concept was mathematically convenient, it was difficult to visualize, and faced opposition.[19] One of its critics, Max Born, proposed instead that Schrodinger's wavefunction described not the electron but rather all its possible states, and thus could be used to calculate the probability of finding an electron at any given location around the nucleus.[20]

In 1927, Werner Heisenberg pointed out that since a wavefunction incorporates time as well as position, it is impossible to simultaneously derive precise values for both the position and momentum of a particle for any given point in time[21]; this became known as the uncertainty principle.

The five atomic orbitals of a neon atom, separated and arranged in order of increasing energy. Each orbital holds up to two electrons, which exist for most of the time in the zones represented by the colored bubbles. Each electron is equally in both orbital zones, shown here by color only to highlight the different wave phase.

This new approach invalidated Bohr's model, with its neat, clearly defined circular orbits. The modern model of the atom describes the positions of electrons in an atom in terms of probabilities. An electron can potentially be found at any distance from the nucleus, but—depending on its energy level—tends to exist more frequently in certain regions around the nucleus than others; this pattern is referred to as its atomic orbital.

See also

Notes

  1. ^ Gangopadhyaya, Mrinalkanti. Indian Atomism: History and Sources. Atlantic Highlands, New Jersey: Humanities Press, 1981. ISBN 0-391-02177-X
  2. ^ "Atom," Microsoft® Encarta® Online Encyclopedia 2007. © 1997-2007 Microsoft Corporation. All Rights Reserved. Retrieved on August 29, 2007.
  3. ^ Gardet, L. “djuz’” in Encyclopaedia of Islam CD-ROM Edition, v. 1.1. Leiden: Brill, 2001.
  4. ^ Weisstein, Eric W. "Lavoisier, Antoine (1743-1794)." scienceworld.wolfram.com. Retrieved on August 29, 2007.
  5. ^ Proust, Joseph Louis. "Researches on Copper," excerpted from Ann. chim. 32, 26-54 (1799) [as translated and reproduced in Henry M. Leicester and Herbert S. Klickstein, A Source Book in Chemistry, 1400-1900 (Cambridge, MA: Harvard, 1952)]. Retrieved on August 29, 2007.
  6. ^ Dalton, John. "On the Absorption of Gases by Water and Other Liquids," in Memoirs of the Literary and Philosophical Society of Manchester. 1803. Retrieved on August 29, 2007.
  7. ^ Johnson, Chris. "Avogadro - his contribution to chemistry." Last modified on July 4, 2004. Retrieved on August 29, 2007.
  8. ^ Avogadro, Amedeo. "Essay on a Manner of Determining the Relative Masses of the Elementary Molecules of Bodies, and the Proportions in Which They Enter into These Compounds." 1811. Journal de Physique, 73, 58-76. Retrieved on August 29, 2007.
  9. ^ Einstein, Albert. "On the Movement of Small Particles Suspended in Stationary Liquids Required by the Molecular-Kinetic Theory of Heat." Annal der Physik. Retrieved on August 29, 2007.
  10. ^ Thomson, J.J. "Cathode rays." Philosophical Magazine, 44, 293 (1897). [facsimile from Stephen Wright, Classical Scientific Papers, Physics (Mills and Boon, 1964).] Retrieved on August 29, 2007.
  11. ^ Thomson, J.J. "On the Structure of the Atom: an Investigation of the Stability and Periods of Oscillation of a number of Corpuscles arranged at equal intervals around the Circumference of a Circle; with Application of the Results to the Theory of Atomic Structure." Philosophical Magazine March 1904. Series 6, Vol 7, No 39. Retrieved on August 29, 2007.
  12. ^ Geiger, H. "The Scattering of the α-Particles by Matter." Proceedings of the Royal Society February 17, 1910. Series A 82: 495–500. Retrieved on August 29, 2007.
  13. ^ Rutherford, Ernest. "The Scattering of α and β Particles by Matter and the Structure of the Atom." Philosophical Magazine. May 1911. Series 6, Vol. 21. Retrieved on August 29, 2007.
  14. ^ Thomson, J.J. "Rays of positive electricity." Proceedings of the Royal Society. 1913. A 89, 1–20 [as excerpted in Henry A. Boorse & Lloyd Motz, The World of the Atom, Vol. 1 (New York: Basic Books, 1966)]. Retrieved on August 29, 2007.
  15. ^ Rutherford, Ernest. "Collisions of alpha Particles with Light Atoms. IV. An Anomalous Effect in Nitrogen." Philosophical Magazine. 1919. 6th series, 37, 581. Retrieved on August 29, 2007.
  16. ^ Chadwick, James. (Feb 27, 1932), "Possible Existence of a Neutron." Nature. February 27, 1932. p. 312. Retrieved on August 29, 2007.
  17. ^ a b c Bohr, N. "On the constitution of atoms and molecules." Philosophical Magazine. July 1913. 26, 1-25. Retrieved on August 29, 2007.
  18. ^ Schrodinger, Erwin. "Quantisation as an Eigenvalue Problem." Annalen der Physik.
  19. ^ Mahanti, Subodh. "Erwin Schrodinger: The Founder of Quantum Wave Mechanics." Retrieved on August 29, 2007.
  20. ^ Mahanti, Subodh. "Max Born: Founder of Lattice Dynamics." Retrieved on August 29, 2007.
  21. ^ "Heisenberg Uncertainty Principle." International Society of Complexity, Information and Design: Encyclopedia. Retrieved on August 29, 2007.