Sei sulla pagina 1di 264

Magnetic Susceptibility

PDF generated using the open source mwlib toolkit. See http://code.pediapress.com/ for more information. PDF generated at: Tue, 01 May 2012 20:09:58 UTC

Contents
Articles
Magnetic Theory
Magnetism Magnetic field Magnetization Magnetic moment Demagnetizing field Magnetic susceptibility Permeability (electromagnetism) Force between magnets Lorentz force 1 1 11 31 33 43 47 51 55 63 74 74 83 84 99 103 104 107 115 119 125 129 131 132 134 135 136 139 139 142 150

Microscopic Origins of Magnetism


Magnetochemistry Unpaired electron Atomic orbital d electron count Hund's rule of maximum multiplicity Aufbau principle Coordination complex Diamagnetism Paramagnetism Electron magnetic dipole moment Pascal's constants CurieWeiss law Curie's law Curie constant Boltzmann distribution Brillouin and Langevin functions

Paramagnetic Resonance
Gyromagnetic ratio Electron paramagnetic resonance Larmor precession

Pulsed EPR

152 156 156 157 168 169 171 177 180 183 186 188 191 192 200 208 209 214 214 220 221 225 230 235 237 242 250 251

Quantum Magnetism
Bohr magneton Spin (physics) Land g-factor g-factor (physics) Zeeman effect Spin states (d electrons) Spinorbit interaction Azimuthal quantum number Principal quantum number Spin quantum number Total angular momentum quantum number Angular momentum operator Angular momentum Magnetic quantum number Pauli exclusion principle

Materials
Ferrofluid Magnetic dipoledipole interaction Magnetic hyperthermia Magnetic nanoparticles Single-molecule magnet Magnetic anisotropy Magnetocrystalline anisotropy Electron configuration Inverse magnetostrictive effect Exchange bias

References
Article Sources and Contributors Image Sources, Licenses and Contributors 253 257

Article Licenses
License 261

Magnetic Theory
Magnetism
Magnetism is a property of materials that respond to an applied magnetic field. Permanent magnets have persistent magnetic fields caused by ferromagnetism. That is the strongest and most familiar type of magnetism. However, all materials are influenced varyingly by the presence of a magnetic field. Some are attracted to a magnetic field (paramagnetism); others are repulsed by a magnetic field (diamagnetism); others have a much more complex relationship with an applied magnetic field (spin glass behavior and antiferromagnetism). Substances that are negligibly affected by magnetic fields are known as non-magnetic substances. They include copper, aluminium, gases, and plastic. Pure oxygen exhibits magnetic properties when cooled to a liquid state. The magnetic state (or phase) of a material depends on temperature (and other variables such as pressure and applied magnetic field) so that a material may exhibit more than one form of magnetism depending on its temperature, etc.

History
Aristotle attributed the first of what could be called a scientific discussion on magnetism to Thales of Miletus, who lived from about 625 BC to about 545 BC.[1] Around the same time, in ancient India, the Indian surgeon, Sushruta, was the first to make use of the magnet for surgical purposes.[2] In ancient China, the earliest literary reference to magnetism lies in a 4th century BC book called Book of the Devil Valley Master ( ): "The lodestone makes iron come or it attracts it."[3] The earliest mention of the attraction of a needle appears in a work composed between AD 20 and 100 (Louen-heng): "A lodestone attracts a needle."[4] The ancient Chinese scientist Shen Kuo (10311095) was the first person to write of the magnetic needle compass and that it improved the accuracy of navigation by employing the astronomical concept of true north (Dream Pool Essays, AD 1088), and by the 12th century the Chinese were known to use the lodestone compass for navigation. They sculpted a directional spoon from lodestone in such a way that the handle of the spoon always pointed south. Alexander Neckham, by 1187, was the first in Europe to describe the compass and its use for navigation. In 1269, Peter Peregrinus de Maricourt wrote the Epistola de magnete, the first extant treatise describing the properties of magnets. In 1282, the properties of magnets and the dry compass were discussed by Al-Ashraf, a Yemeni physicist, astronomer, and geographer.[5]

Magnetism

In 1600, William Gilbert published his De Magnete, Magneticisque Corporibus, et de Magno Magnete Tellure (On the Magnet and Magnetic Bodies, and on the Great Magnet the Earth). In this work he describes many of his experiments with his model earth called the terrella. From his experiments, he concluded that the Earth was itself magnetic and that this was the reason compasses pointed north (previously, some believed that it was the pole star (Polaris) or a large magnetic island on the north pole that attracted the compass). An understanding of the relationship between electricity and magnetism began in 1819 with work by Hans Christian Oersted, a professor at the University of Copenhagen, who discovered more or less by accident that an electric current could influence a compass needle. This landmark experiment is known as Oersted's Experiment. Several other experiments Michael Faraday, 1842 followed, with Andr-Marie Ampre, who in 1820 discovered that the magnetic field circulating in a closed-path was related to the current flowing through the perimeter of the path; Carl Friedrich Gauss; Jean-Baptiste Biot and Flix Savart, both of which in 1820 came up with the Biot-Savart Law giving an equation for the magnetic field from a current-carrying wire; Michael Faraday, who in 1831 found that a time-varying magnetic flux through a loop of wire induced a voltage, and others finding further links between magnetism and electricity. James Clerk Maxwell synthesized and expanded these insights into Maxwell's equations, unifying electricity, magnetism, and optics into the field of electromagnetism. In 1905, Einstein used these laws in motivating his theory of special relativity,[6] requiring that the laws held true in all inertial reference frames. Electromagnetism has continued to develop into the 21st century, being incorporated into the more fundamental theories of gauge theory, quantum electrodynamics, electroweak theory, and finally the standard model.

Sources of magnetism
Magnetism, at its root, arises from two sources: 1. Electric currents or more generally, moving electric charges create magnetic fields (see Maxwell's Equations). 2. Many particles have nonzero "intrinsic" (or "spin") magnetic moments. Just as each particle, by its nature, has a certain mass and charge, each has a certain magnetic moment, possibly zero. It was found hundreds of years ago that certain materials have a tendency to orient in a particular direction. For example ancient people knew that "lodestones," when suspended from a string and allowed to freely rotate, come to rest horizontally in the North-South direction. Ancient Mariners used lodestones for navigational purposes. In magnetic materials, sources of magnetization are the electrons' orbital angular motion around the nucleus, and the electrons' intrinsic magnetic moment (see electron magnetic dipole moment). The other sources of magnetism are the nuclear magnetic moments of the nuclei in the material which are typically thousands of times smaller than the electrons' magnetic moments, so they are negligible in the context of the magnetization of materials. Nuclear magnetic moments are important in other contexts, particularly in nuclear magnetic resonance (NMR) and magnetic resonance imaging (MRI). Ordinarily, the enormous number of electrons in a material are arranged such that their magnetic moments (both orbital and intrinsic) cancel out. This is due, to some extent, to electrons combining into pairs with opposite intrinsic magnetic moments as a result of the Pauli exclusion principle (see electron configuration), or combining into filled subshells with zero net orbital motion. In both cases, the electron arrangement is so as to exactly cancel the magnetic moments from each electron. Moreover, even when the electron configuration is such that there are unpaired electrons and/or non-filled subshells, it is often the case that the various electrons in the solid will contribute

Magnetism magnetic moments that point in different, random directions, so that the material will not be magnetic. However, sometimes either spontaneously, or owing to an applied external magnetic field each of the electron magnetic moments will be, on average, lined up. Then the material can produce a net total magnetic field, which can potentially be quite strong. The magnetic behavior of a material depends on its structure, particularly its electron configuration, for the reasons mentioned above, and also on the temperature. At high temperatures, random thermal motion makes it more difficult for the electrons to maintain alignment.

Topics
Diamagnetism
Diamagnetism appears in all materials, and is the tendency of a material to oppose an applied magnetic field, and therefore, to be repelled by a magnetic field. However, in a material with paramagnetic properties (that is, with a tendency to enhance an external magnetic field), the paramagnetic behavior dominates.[8] Thus, despite its universal occurrence, diamagnetic behavior is observed only in a purely [7] Hierarchy of types of magnetism. diamagnetic material. In a diamagnetic material, there are no unpaired electrons, so the intrinsic electron magnetic moments cannot produce any bulk effect. In these cases, the magnetization arises from the electrons' orbital motions, which can be understood classically as follows: When a material is put in a magnetic field, the electrons circling the nucleus will experience, in addition to their Coulomb attraction to the nucleus, a Lorentz force from the magnetic field. Depending on which direction the electron is orbiting, this force may increase the centripetal force on the electrons, pulling them in towards the nucleus, or it may decrease the force, pulling them away from the nucleus. This effect systematically increases the orbital magnetic moments that were aligned opposite the field, and decreases the ones aligned parallel to the field (in accordance with Lenz's law). This results in a small bulk magnetic moment, with an opposite direction to the applied field. Note that this description is meant only as an heuristic; a proper understanding requires a quantum-mechanical description. Note that all materials undergo this orbital response. However, in paramagnetic and ferromagnetic substances, the diamagnetic effect is overwhelmed by the much stronger effects caused by the unpaired electrons.

Paramagnetism
In a paramagnetic material there are unpaired electrons, i.e. atomic or molecular orbitals with exactly one electron in them. While paired electrons are required by the Pauli exclusion principle to have their intrinsic ('spin') magnetic moments pointing in opposite directions, causing their magnetic fields to cancel out, an unpaired electron is free to align its magnetic moment in any direction. When an external magnetic field is applied, these magnetic moments will tend to align themselves in the same direction as the applied field, thus reinforcing it.

Magnetism

Ferromagnetism
A ferromagnet, like a paramagnetic substance, has unpaired electrons. However, in addition to the electrons' intrinsic magnetic moment's tendency to be parallel to an applied field, there is also in these materials a tendency for these magnetic moments to orient parallel to each other to maintain a lowered-energy state. Thus, even when the applied field is removed, the electrons in the material maintain a parallel orientation. Every ferromagnetic substance has its own individual temperature, called the Curie temperature, or Curie point, above which it loses its ferromagnetic properties. This is because the thermal tendency to disorder overwhelms the energy-lowering due to ferromagnetic order.

Permanent magnet holding up some coins

Some well-known ferromagnetic materials that exhibit easily detectable magnetic properties (to form magnets) are nickel, iron, cobalt, gadolinium and their alloys. Magnetic domains The magnetic moment of atoms in a ferromagnetic material cause them to behave something like tiny permanent magnets. They stick together and align themselves into small regions of more or less uniform alignment called magnetic domains or Weiss domains. Magnetic domains can be observed with a magnetic force microscope to reveal magnetic domain boundaries that resemble white lines in the sketch. There are many scientific experiments that can physically show magnetic fields.

Magnetic domains in ferromagnetic material.

When a domain contains too many molecules, it becomes unstable and divides into two domains aligned in opposite directions so that they stick together more stably as shown at the right. When exposed to a magnetic field, the domain boundaries move so that the domains aligned with the magnetic field grow and dominate the structure as shown at the left. When the magnetizing field is removed, the domains may not return to an unmagnetized state. This results in the ferromagnetic material's being magnetized, forming a permanent magnet. When magnetized strongly enough that the prevailing domain overruns all others to result in only one single domain, the material is magnetically saturated. When a magnetized ferromagnetic material is heated to the Curie point temperature, the molecules are
Effect of a magnet on the domains.

Magnetism agitated to the point that the magnetic domains lose the organization and the magnetic properties they cause cease. When the material is cooled, this domain alignment structure spontaneously returns, in a manner roughly analogous to how a liquid can freeze into a crystalline solid.

Antiferromagnetism
In an antiferromagnet, unlike a ferromagnet, there is a tendency for the intrinsic magnetic moments of neighboring valence electrons to point in opposite directions. When all atoms are arranged in a substance so that each neighbor is 'anti-aligned', the substance is antiferromagnetic. Antiferromagnets have a zero net magnetic moment, meaning no field is produced by them. Antiferromagnets are less common compared to the other types of behaviors, and are mostly observed at low temperatures. In varying temperatures, antiferromagnets can be seen to exhibit diamagnetic and ferrimagnetic properties.

Antiferromagnetic ordering

In some materials, neighboring electrons want to point in opposite directions, but there is no geometrical arrangement in which each pair of neighbors is anti-aligned. This is called a spin glass, and is an example of geometrical frustration.

Ferrimagnetism
Like ferromagnetism, ferrimagnets retain their magnetization in the absence of a field. However, like antiferromagnets, neighboring pairs of electron spins like to point in opposite directions. These two properties are not contradictory, because in the optimal geometrical arrangement, there is more magnetic moment from the sublattice of electrons that point in one direction, than from the sublattice that points in the opposite direction.

Ferrimagnetic ordering

The first discovered magnetic substance, magnetite, was originally believed to be a ferromagnet; Louis Nel disproved this, however, with the discovery of ferrimagnetism.

Magnetism

Superparamagnetism
When a ferromagnet or ferrimagnet is sufficiently small, it acts like a single magnetic spin that is subject to Brownian motion. Its response to a magnetic field is qualitatively similar to the response of a paramagnet, but much larger.

Electromagnet
An electromagnet is a type of magnet whose magnetism is produced by the flow of electric current. The magnetic field disappears when the current ceases.

Other types of magnetism


Molecular magnet Metamagnetism Molecule-based magnet Spin glass

Magnetism, electricity, and special relativity


As a consequence of Einstein's theory of special relativity, electricity and magnetism are fundamentally interlinked. Both magnetism lacking electricity, and electricity without magnetism, are inconsistent with special relativity, due to such effects as length contraction, time dilation, and the fact that the magnetic force is velocity-dependent. However, when both electricity and magnetism are taken into account, the resulting theory (electromagnetism) is fully consistent with special relativity.[6][9] In particular, a phenomenon that appears purely electric to one observer may be purely magnetic to another, or more generally the relative contributions of electricity and magnetism are dependent on the frame of reference. Thus, special relativity "mixes" electricity and magnetism into a single, inseparable phenomenon called electromagnetism, analogous to how relativity "mixes" space and time into spacetime.
Electromagnets attracts paper clips when current is applied creating a magnetic field. The electromagnet loses them when current and magnetic field are removed.

Magnetic fields in a material


In a vacuum,

where 0 is the vacuum permeability. In a material,

The quantity 0M is called magnetic polarization. If the field H is small, the response of the magnetization M in a diamagnet or paramagnet is approximately linear:

the constant of proportionality being called the magnetic susceptibility. If so,

In a hard magnet such as a ferromagnet, M is not proportional to the field and is generally nonzero even when H is zero (see Remanence).

Magnetism

Force due to magnetic field - The magnetic force


The phenomenon of magnetism is "mediated" by the magnetic field. An electric current or magnetic dipole creates a magnetic field, and that field, in turn, imparts magnetic forces on other particles that are in the fields. Maxwell's equations, which simplify to the Biot-Savart law in the case of steady currents, describe the origin and behavior of the fields that govern these forces. Therefore magnetism is seen whenever electrically charged particles are in motion---for example, from movement of electrons in an electric current, or in certain cases from the orbital motion of electrons around an atom's nucleus. They also arise from "intrinsic" magnetic dipoles arising from quantum-mechanical spin.

Magnetic lines of force of a bar magnet shown by iron filings on paper

The same situations that create magnetic fields charge moving in a current or in an atom, and intrinsic magnetic dipoles are also the situations in which a magnetic field has an effect, creating a force. Following is the formula for moving charge; for the forces on an intrinsic dipole, see magnetic dipole. When a charged particle moves through a magnetic field B, it feels a Lorentz force F given by the cross product:[10]

where is the electric charge of the particle, and v is the velocity vector of the particle Because this is a cross product, the force is perpendicular to both the motion of the particle and the magnetic field. It follows that the magnetic force does no work on the particle; it may change the direction of the particle's movement, but it cannot cause it to speed up or slow down. The magnitude of the force is

where

is the angle between v and B.

One tool for determining the direction of the velocity vector of a moving charge, the magnetic field, and the force exerted is labeling the index finger "V", the middle finger "B", and the thumb "F" with your right hand. When making a gun-like configuration, with the middle finger crossing under the index finger, the fingers represent the velocity vector, magnetic field vector, and force vector, respectively. See also right hand rule.

Magnetic dipoles
A very common source of magnetic field shown in nature is a dipole, with a "South pole" and a "North pole", terms dating back to the use of magnets as compasses, interacting with the Earth's magnetic field to indicate North and South on the globe. Since opposite ends of magnets are attracted, the north pole of a magnet is attracted to the south pole of another magnet. The Earth's North Magnetic Pole (currently in the Arctic Ocean, north of Canada) is physically a south pole, as it attracts the north pole of a compass. A magnetic field contains energy, and physical systems move toward configurations with lower energy. When diamagnetic material is placed in a magnetic field, a magnetic dipole tends to align itself in opposed polarity to that field, thereby lowering the net field strength. When ferromagnetic material is placed within a magnetic field, the magnetic dipoles align to the applied field, thus expanding the domain walls of the magnetic domains.

Magnetism

Magnetic monopoles
Since a bar magnet gets its ferromagnetism from electrons distributed evenly throughout the bar, when a bar magnet is cut in half, each of the resulting pieces is a smaller bar magnet. Even though a magnet is said to have a north pole and a south pole, these two poles cannot be separated from each other. A monopole if such a thing exists would be a new and fundamentally different kind of magnetic object. It would act as an isolated north pole, not attached to a south pole, or vice versa. Monopoles would carry "magnetic charge" analogous to electric charge. Despite systematic searches since 1931, as of 2010, they have never been observed, and could very well not exist.[11] Nevertheless, some theoretical physics models predict the existence of these magnetic monopoles. Paul Dirac observed in 1931 that, because electricity and magnetism show a certain symmetry, just as quantum theory predicts that individual positive or negative electric charges can be observed without the opposing charge, isolated South or North magnetic poles should be observable. Using quantum theory Dirac showed that if magnetic monopoles exist, then one could explain the quantization of electric charge---that is, why the observed elementary particles carry charges that are multiples of the charge of the electron. Certain grand unified theories predict the existence of monopoles which, unlike elementary particles, are solitons (localized energy packets). The initial results of using these models to estimate the number of monopoles created in the big bang contradicted cosmological observations the monopoles would have been so plentiful and massive that they would have long since halted the expansion of the universe. However, the idea of inflation (for which this problem served as a partial motivation) was successful in solving this problem, creating models in which monopoles existed but were rare enough to be consistent with current observations.[12]

Quantum-mechanical origin of magnetism


In principle all kinds of magnetism originate (similar to Superconductivity) from specific quantum-mechanical phenomena (e.g. Mathematical formulation of quantum mechanics, in particular the chapters on spin and on the Pauli principle). A successful model was developed already in 1927, by Walter Heitler and Fritz London, who derived quantum-mechanically, how hydrogen molecules are formed from hydrogen atoms, i.e. from the atomic hydrogen orbitals and centered at the nuclei A and B, see below. That this leads to magnetism, is not at all obvious, but will be explained in the following. According the Heitler-London theory, so-called two-body molecular orbital is: -orbitals are formed, namely the resulting

Here the last product means that a first electron, r1, is in an atomic hydrogen-orbital centered at the second nucleus, whereas the second electron runs around the first nucleus. This "exchange" phenomenon is an expression for the quantum-mechanical property that particles with identical properties cannot be distinguished. It is specific not only for the formation of chemical bonds, but as we will see, also for magnetism, i.e. in this connection the term exchange interaction arises, a term which is essential for the origin of magnetism, and which is stronger, roughly by factors 100 and even by 1000, than the energies arising from the electrodynamic dipole-dipole interaction. As for the spin function , which is responsible for the magnetism, we have the already mentioned Pauli's

principle, namely that a symmetric orbital (i.e. with the + sign as above) must be multiplied with an antisymmetric spin function (i.e. with a - sign), and vice versa. Thus: , I.e., not only and must be substituted by and , respectively (the first entity means "spin up", the second one "spin down"), but also the sign + by the sign, and finally ri by the discrete values si (=); thereby we have and . The "singlet state", i.e. the - sign, means: the

Magnetism spins are antiparallel, i.e. for the solid we have antiferromagnetism, and for two-atomic molecules one has diamagnetism. The tendency to form a (homoeopolar) chemical bond (this means: the formation of a symmetric molecular orbital, i.e. with the + sign) results through the Pauli principle automatically in an antisymmetric spin state (i.e. with the - sign). In contrast, the Coulomb repulsion of the electrons, i.e. the tendency that they try to avoid each other by this repulsion, would lead to an antisymmetric orbital function (i.e. with the - sign) of these two particles, and complementary to a symmetric spin function (i.e. with the + sign, one of the so-called "triplet functions"). Thus, now the spins would be parallel (ferromagnetism in a solid, paramagnetism in two-atomic gases). The last-mentioned tendency dominates in the metals iron, cobalt and nickel, and in some rare earths, which are ferromagnetic. Most of the other metals, where the first-mentioned tendency dominates, are nonmagnetic (e.g. sodium, aluminium, and magnesium) or antiferromagnetic (e.g. manganese). Diatomic gases are also almost exclusively diamagnetic, and not paramagnetic. However, the oxygen molecule, because of the involvement of -orbitals, is an exception important for the life-sciences. The Heitler-London considerations can be generalized to the Heisenberg model of magnetism (Heisenberg 1928). The explanation of the phenomena is thus essentially based on all subtleties of quantum mechanics, whereas the electrodynamics covers mainly the phenomenology.

Units of electromagnetism
SI units related to magnetism
SI electromagnetism units Symbol [13] Name of Quantity Electric current Electric charge Potential difference; Electromotive force Derived Units ampere (SI base unit) coulomb volt Conversion of International to SI base units

Electric resistance; Impedance; Reactance ohm Resistivity Electric power Capacitance Electric field strength Electric displacement field Permittivity Electric susceptibility Conductance; Admittance; Susceptance Conductivity ohm metre watt farad volt per metre Coulomb per square metre farad per metre Dimensionless siemens siemens per metre

Magnetic flux density, Magnetic induction tesla Magnetic flux Magnetic field strength Inductance Permeability Magnetic susceptibility weber ampere per metre henry henry per metre Dimensionless

Magnetism

10

Other units
gauss The gauss, abbreviated as G, is the CGS unit of magnetic field (B). oersted The oersted is the CGS unit of magnetizing field (H). Maxwell is the CGS unit for the magnetic flux. gamma is a unit of magnetic flux density that was commonly used before the tesla became popular (1 gamma = 1 nT) 0 common symbol for the permeability of free space (4107 N/(ampere-turn)2).

Living things
Some organisms can detect magnetic fields, a phenomenon known as magnetoception. Magnetobiology studies magnetic fields as a medical treatment; fields naturally produced by an organism are known as biomagnetism.

References
[1] Fowler, Michael (1997). "Historical Beginnings of Theories of Electricity and Magnetism" (http:/ / galileoandeinstein. physics. virginia. edu/ more_stuff/ E& M_Hist. html). . Retrieved 2008-04-02. [2] Vowles, Hugh P. (1932). "Early Evolution of Power Engineering". Isis (University of Chicago Press) 17 (2): 412420 [41920]. doi:10.1086/346662. [3] [4] [5] [6] [7] Li Shu-hua, Origine de la Boussole 11. Aimant et Boussole, Isis, Vol. 45, No. 2. (Jul., 1954), p.175 Li Shu-hua, Origine de la Boussole 11. Aimant et Boussole, Isis, Vol. 45, No. 2. (Jul., 1954), p.176 Schmidl, Petra G. (19961997). "Two Early Arabic Sources On The Magnetic Compass". Journal of Arabic and Islamic Studies 1: 81132. A. Einstein: "On the Electrodynamics of Moving Bodies" (http:/ / www. fourmilab. ch/ etexts/ einstein/ specrel/ www/ ), June 30, 1905. HP Meyers (1997). Introductory solid state physics (http:/ / books. google. com/ ?id=Uc1pCo5TrYUC& pg=PA322) (2 ed.). CRC Press. p.362; Figure 11.1. ISBN0748406603. . [8] Catherine Westbrook, Carolyn Kaut, Carolyn Kaut-Roth (1998). MRI (Magnetic Resonance Imaging) in practice (http:/ / books. google. com/ ?id=Qq1SHDtS2G8C& pg=PA217) (2 ed.). Wiley-Blackwell. p.217. ISBN0632042052. . [9] Griffiths 1998, chapter 12 [10] Jackson, John David (1999). Classical electrodynamics (3rd ed.). New York, [NY.]: Wiley. ISBN0-471-30932-X [11] Milton mentions some inconclusive events (p.60) and still concludes that "no evidence at all of magnetic monopoles has survived" (p.3). Milton, Kimball A. (June 2006). "Theoretical and experimental status of magnetic monopoles". Reports on Progress in Physics 69 (6): 16371711. arXiv:hep-ex/0602040. Bibcode2006RPPh...69.1637M. doi:10.1088/0034-4885/69/6/R02.. [12] Guth, Alan (1997). The Inflationary Universe: The Quest for a New Theory of Cosmic Origins. Perseus. ISBN0-201-32840-2. OCLC38941224.. [13] International Union of Pure and Applied Chemistry (1993). Quantities, Units and Symbols in Physical Chemistry, 2nd edition, Oxford: Blackwell Science. ISBN 0-632-03583-8. pp.1415. Electronic version. (http:/ / old. iupac. org/ publications/ books/ gbook/ green_book_2ed. pdf)

Further reading
Furlani, Edward P. (2001). Permanent Magnet and Electromechanical Devices: Materials, Analysis and Applications. Academic Press. ISBN0-12-269951-3. OCLC162129430. Griffiths, David J. (1998). Introduction to Electrodynamics (3rd ed.). Prentice Hall. ISBN0-13-805326-X. OCLC40251748. Kronmller, Helmut. (2007). Handbook of Magnetism and Advanced Magnetic Materials, 5 Volume Set. John Wiley & Sons. ISBN978-0-470-02217-7. OCLC124165851. Tipler, Paul (2004). Physics for Scientists and Engineers: Electricity, Magnetism, Light, and Elementary Modern Physics (5th ed.). W. H. Freeman. ISBN0-7167-0810-8. OCLC51095685. David K. Cheng (1992). Field and Wave Electromagnetics. Addison-Wesley Publishing Company, Inc.. ISBN0-201-12819-5.

Magnetism

11

External links
Magnetism (http://www.bbc.co.uk/programmes/p003k9dd) on In Our Time at the BBC. ( listen now (http:// www.bbc.co.uk/iplayer/console/p003k9dd/In_Our_Time_Magnetism)) The Exploratorium Science Snacks Snacks about Magnetism (http://www.exploratorium.edu/snacks/ iconmagnetism.html) Electromagnetism (http://www.lightandmatter.com/html_books/0sn/ch11/ch11.html) - a chapter from an online textbook Video: The physicist Richard Feynman answers the question, Why do bar magnets attract or repel each other? (http://www.youtube.com/watch?v=wMFPe-DwULM) On the Magnet, 1600 (http://www.antiquebooks.net/readpage.html#gilbert) First scientific book on magnetism by the father of electrical engineering. Full English text, full text search.

Magnetic field
A magnetic field is a mathematical description of the magnetic influence of electric currents and magnetic materials. The magnetic field at any given point is specified by both a direction and a magnitude (or strength); as such it is a vector field.[1] The magnetic field is most commonly defined in terms of the Lorentz force it exerts on moving electric charges. There are two separate but closely related fields to which the name 'magnetic field' can refer: a magnetic B field and a magnetic H field. Magnetic fields are produced by moving electric charges and the intrinsic magnetic moments of elementary particles associated with a fundamental quantum property, their spin. In special relativity, electric and magnetic fields are two interrelated aspects of a single object, called the electromagnetic field tensor; the aspect of the electromagnetic field that is seen as a magnetic field is dependent on the reference frame of the observer. In quantum physics, the electromagnetic field is quantized and electromagnetic interactions result from the exchange of photons. Magnetic fields have had many uses in ancient and modern society. The Earth produces its own magnetic field, which is important in navigation. Rotating magnetic fields are utilized in both electric motors and generators. Magnetic forces give information about the charge carriers in a material through the Hall effect. The interaction of magnetic fields in electric devices such as transformers is studied in the discipline of magnetic circuits.

Magnetic field

12

History
Although magnets and magnetism were known much earlier, one of the first descriptions of the magnetic field was produced in 1269 by the French scholar Petrus Peregrinus[2] who mapped out the magnetic field on the surface of a spherical magnet using iron needles. Noting that the resulting field lines crossed at two points he named those points 'poles' in analogy to Earth's poles. Almost three centuries later, William Gilbert of Colchester replicated Petrus Peregrinus' work and was the first to state explicitly that Earth itself was a magnet.[3] Gilbert's great work De Magnete was published in 1600 and helped to establish the study of magnetism as a science.

In 1750, John Michell published[4] that magnetic poles attract and repel with an inverse square law.[5] Later, in 1785, Charles-Augustin de Coulomb verified Michell's law and stated explicitly that the North and South poles cannot be separated.[6] One of the first successful models of the magnetic field was developed in 1824 by Simon-Denis Poisson (17811840).[7] His model was built on the inverse square law force between magnetic poles and is exactly analogous to modern electrostatics with a magnetic H-field being produced by 'magnetic poles' in the same way that an electric field E-field is produced by electric charges. Poisson's model also introduced the magnetic scalar potential. Three remarkable discoveries though, would challenge Poisson's model. First, in 1819, Hans Christian Oersted discovered that an electric current generates a magnetic field encircling it. Then, Andr-Marie Ampre showed that parallel wires having currents in the same direction attract one another. Finally Jean-Baptiste Biot and Flix Savart discovered the BiotSavart law which correctly predicts the magnetic field around any current-carrying wire. In 1825, Ampere published the results of his comprehensive study of the magnetic properties of currents for which he coined the term electrodynamics. In it he showed the equivalence of electrical currents to magnets.[8] With this equivalence, Ampre viewed the current loops as being the more fundamental and imagined that magnetic molecules contained perpetually flowing loops of current. In this model, later called the Amprian loop model, current loops (called magnetic dipoles) would replace the dipoles of charge of the Poisson's model.[9] No magnetic charges are needed which has the additional benefit of explaining why magnetic charge can not be isolated. In this work as well, Ampre derived his Ampre's force law which describes the force between two currents. Finally it included his Ampre's law which like the BiotSavart law correctly described the magnetic field generated by a steady current but was more general. In 1831, Michael Faraday showed that a changing magnetic field generates an encircling electric field and thereby demonstrated that electricity and magnetism are even more tightly knitted. Between 1861 and 1865, James Clerk Maxwell developed and published a set of Maxwell's equations which explained and united all of classical electricity and magnetism. The first set of these equations was published in a paper entitled On Physical Lines of Force in 1861. The mechanism that Maxwell proposed to underlie these

One of the first drawings of a magnetic field, by Ren Descartes, 1644. It illustrated his theory that magnetism was caused by the circulation of tiny helical particles, "threaded parts", through threaded pores in magnets.

Magnetic field equations in this paper was fundamentally incorrect, which is not surprising since it predated the modern understanding even of the atom. Yet, the equations were valid although incomplete. He completed the set of Maxwell's equations in his later 1865 paper A Dynamical Theory of the Electromagnetic Field and demonstrated the fact that light is an electromagnetic wave. Thus, he theoretically unified not only electricity and magnetism but light as well. This fact was then later confirmed experimentally by Heinrich Hertz in 1887. Even though the classical theory of electrodynamics was essentially complete with Maxwell's equations, the twentieth century saw a number of improvements and extensions to the theory. Albert Einstein, in his great paper of 1905 that established relativity, showed that both the electric and magnetic fields are part of the same phenomena viewed from different reference frames. (See moving magnet and conductor problem for details about the thought experiment that eventually helped Albert Einstein to develop special relativity.) Finally, the emergent field of quantum mechanics was merged with electrodynamics to form quantum electrodynamics or QED.

13

Definitions, units, and measurement


Magnetic field can be defined in many equivalent ways based on the effects it has on its environment. For instance, a particle having an electric charge, q, and moving in a magnetic field with a velocity, v, experiences a force, F, called the Lorentz force. See force on a charged particle below. Alternatively, the magnetic field can be defined in terms of the torque it produces on a magnetic dipole. See magnetic torque on permanent magnets below. Devices used to measure the local magnetic field are called magnetometers. Important classes of magnetometers include using a rotating coil, Hall effect magnetometers, NMR magnetometers, SQUID magnetometers, and fluxgate magnetometers. The magnetic fields of distant astronomical objects are measured through their effects on local charged particles. For instance, electrons spiraling around a field line produce synchrotron radiation which is detectable in radio waves.
Alternative names for B Magnetic flux density Magnetic induction Magnetic field [10][11] [10]

Alternative names for H

Magnetic field intensity Magnetic field strength Magnetic field Magnetizing field

There are two magnetic fields, H and B. In a vacuum they are indistinguishable, differing only by a multiplicative constant that depends on the physical units. Inside a material they are different (see H and B inside and outside of magnetic materials). The term magnetic field is historically reserved for H while using other terms for B. Informally, though, and formally for some recent textbooks mostly in physics, the term 'magnetic field' is used to describe B as well as or in place of H.[12] There are many alternative names for both (see sidebar to right). The B-field is measured in teslas in SI units and in gauss in cgs units. (1 tesla = 10,000 gauss). The SI unit of tesla is equivalent to (newtonsecond)/(coulombmetre).[13] The H-field is measured in ampere-turn per metre (A/m) in SI units, and in oersteds (Oe) in cgs units.[14] The smallest precision level for a magnetic field measurement[15] is on the order of attoteslas (1018 tesla); the largest magnetic field produced in a laboratory is 2.8kT (VNIIEF in Sarov, Russia, 1998).[16] The magnetic field of some astronomical objects such as magnetars are much higher; magnetars range from 0.1 to 100GT (108 to 1011T).[17] See orders of magnitude (magnetic field).

Magnetic field

14

Magnetic field lines


Mapping the magnetic field of an object is simple in principle. First, measure the strength and direction of the magnetic field at a large number of locations. Then, mark each location with an arrow (called a vector) pointing in the direction of the local magnetic field with a length proportional to the strength of the magnetic field. A simpler way to visualize the magnetic field is to 'connect' the arrows to form magnetic field lines. Magnetic field lines make it much easier to visualize and understand the complex mathematical relationships underlying magnetic field. If done carefully, a field line diagram contains the same information as the vector field it represents. The magnetic field can be estimated at any point on a magnetic field line diagram (whether on a field line or not) using the direction and density of nearby magnetic field lines.[18] A higher density of nearby field lines indicates a stronger magnetic field.
Compasses reveal the direction of the local magnetic field. As seen here, the magnetic field points towards a magnet's south pole and away from its north pole.

Various phenomena have the effect of "displaying" magnetic field lines as though the field lines are physical phenomena. For example, iron filings placed in a magnetic field line up to form lines that correspond to 'field lines'. Magnetic fields "lines" are also visually displayed in polar auroras, in which plasma particle dipole interactions create visible streaks of light that line up with the local direction of Earth's magnetic field. However, field lines are a visual and conceptual aid only and are no more real than (for example) the contour lines (constant altitude) on a topographic map. They do not exist in the actual field; a different choice of mapping scale could show twice as many "lines" or half as many.

Field lines can be used as a qualitative tool to visualize magnetic forces. In ferromagnetic substances like iron and in plasmas, magnetic forces can be understood by imagining that the field lines exert a tension, (like a rubber band) along their length, and a pressure perpendicular to their length on neighboring field lines. 'Unlike' poles of magnets attract because they are linked by many field lines; 'like' poles repel because their field lines do not meet, but run parallel, pushing on each other. Most physical phenomena that "display" magnetic field lines do not include which direction along the lines that the magnetic field is in. A compass, though, reveals that magnetic field lines outside of a magnet point from the north pole (compass points away from north pole) to the south (compass points toward the south pole). The magnetic field of a straight current-carrying wire encircles the wire with a direction that depends on the direction of the current and that can be measured with a compass as well.

The direction of magnetic field lines represented by the alignment of iron filings sprinkled on paper placed above a bar magnet. The mutual attraction of opposite poles of the iron filings results in the formation of elongated clusters of filings along "field lines". The field is not precisely the same as around the isolated magnet; the magnetization of the filings alters the field somewhat.

Magnetic field

15

Magnetic field and permanent magnets


Permanent magnets are objects that produce their own persistent magnetic fields. They are made of ferromagnetic materials, such as iron and nickel, that have been magnetized, and they have both a north and a south pole.

Magnetic field of permanent magnets


The magnetic field of permanent magnets can be quite complicated, especially near the magnet. The magnetic field of a small[19] straight magnet is proportional to the magnet's strength (called its magnetic dipole moment m). The equations are non-trivial and also depend on the distance from the magnet and the orientation of the magnet. For simple magnets, m points in the direction of a line drawn from the south to the north pole of the magnet. Flipping a bar magnet is equivalent to rotating its m by 180 degrees. The magnetic field of larger magnets can be obtained by modelling them as a collection of a large number of small magnets called dipoles each having their own m. The magnetic field produced by the magnet then is the net magnetic field of these dipoles. And, any net force on the magnet is a result of adding up the forces on the individual dipoles. There are two competing models for the nature of these dipoles. These two models produce two different magnetic fields, H and B. Outside a material, though, the two are identical (to a multiplicative constant) so that in many cases the distinction can be ignored. This is particularly true for magnetic fields, such as those due to electric currents, that are not generated by magnetic materials.

Magnetic pole model and the H-field


It is sometimes useful to model the force and torques between two magnets as due to magnetic poles repelling or attracting each other in the same manner as the Coulomb force between electric charges. In this model, a magnetic H-field is produced by magnetic charges that are 'smeared' around each pole. The H-field, therefore, is analogous to the electric field E which starts at a positive electric charge and ends at a negative electric charge. Near the north pole, therefore, all H-field lines point away from the north pole (whether inside the magnet or out) while near the south pole (whether inside the magnet or out) all H-field lines point toward the south pole. A north pole, then, feels a force in the direction of the H-field while the force on the south pole is opposite to the H-field. In the magnetic pole model, the elementary magnetic dipole m is formed by two opposite magnetic poles of pole strength qm separated by a very small distance d, such that m = qm d.

The magnetic pole model: two opposing poles, North (+) and South (-), separated by a distance d produce an H-field (lines).

Unfortunately, magnetic poles cannot exist apart from each other; all magnets have north/south pairs which cannot be separated without creating two magnets each having a north/south pair. Further, magnetic poles do not account for magnetism that is produced by electric currents nor the force that a magnetic field applies to moving electric charges.

Magnetic field

16

Amperian loop model and the B-field


After Oersted discovered that electric currents produce a magnetic field and Ampere discovered that electric currents attracted and repelled each other similar to magnets, it was natural to hypothesize that all magnetic fields are due to electric current loops. In this model developed by Ampere, the elementary magnetic dipole that makes up all magnets is a sufficiently small Amperian loop of current I. The dipole moment of this loop is m = I A where A is the area of the loop. These magnetic dipoles produce a magnetic B field. One important property of the B-field produced this way is that magnetic B field lines neither start nor end (mathematically, B is a solenoidal vector field); a field line either extends to infinity or wraps around to form a closed curve.[20] To date no exception to this rule has been The Amperian loop model: A current loop (ring) which found. (See magnetic monopole below.) Magnetic field lines exit a goes into the page at the x and comes out at the dot produces a B field (lines). The north pole is to the right magnet near its north pole and enter near its south pole, but inside and the south to the left. the magnet B-field lines continue through the magnet from the [21] south pole back to the north. If a B-field line enters a magnet somewhere it has to leave somewhere else; it is not allowed to have an end point. Magnetic poles, therefore, always come in N and S pairs. More formally, since all the magnetic field lines that enter any given region must also leave that region, subtracting the 'number'[22] of field lines that enter the region from the number that exit gives identically zero. Mathematically this is equivalent to: , where the integral is a surface integral over the closed surface S (a closed surface is one that completely surrounds a region with no holes to let any field lines escape). Since dA points outward, the dot product in the integral is positive for B-field pointing out and negative for B-field pointing in. There is also a corresponding differential form of this equation covered in Maxwell's equations below.

Force between magnets


The force between two small magnets is quite complicated and depends on the strength and orientation of both magnets and the distance and direction of the magnets relative to each other. The force is particularly sensitive to rotations of the magnets due to magnetic torque. The force on each magnet depends on its magnetic moment and the magnetic field[23] of the other. To understand the force between magnets, it is useful to examine the magnetic pole model given above. In this model, the H-field of one magnet pushes and pulls on both poles of a second magnet. If this H-field is the same at both poles of the second magnet then there is no net force on that magnet since the force is opposite for opposite poles. If, however, the magnetic field of the first magnet is nonuniform (such as the H near one of its poles), each pole of the second magnet sees a different field and is subject to a different force. This difference in the two forces moves the magnet in the direction of increasing magnetic field and may also cause a net torque. This is a specific example of a general rule that magnets are attracted (or repulsed depending on the orientation of the magnet) into regions of higher magnetic field. Any non-uniform magnetic field whether caused by permanent magnets or by electric currents will exert a force on a small magnet in this way.

Magnetic field The details of the Amperian loop model are different and more complicated but yield the same result: that magnetic dipoles are attracted/repelled into regions of higher magnetic field. Mathematically, the force on a small magnet having a magnetic moment m due to a magnetic field B is:[24]

17

where the gradient is the change of the quantity mB per unit distance and the direction is that of maximum increase of mB. To understand this equation, note that the dot product mB = mBcos(), where m and B represent the magnitude of the m and B vectors and is the angle between them. If m is in the same direction as B then the dot product is positive and the gradient points 'uphill' pulling the magnet into regions of higher B-field (more strictly larger mB). This equation is strictly only valid for magnets of zero size, but is often a good approximation for not too large magnets. The magnetic force on larger magnets is determined by dividing them into smaller regions having their own m then summing up the forces on each of these regions.

Magnetic torque on permanent magnets


If two like poles of two separate magnets are brought near each other and one of the magnets is allowed to turn it will promptly rotate to align itself with the first. In this example, the magnetic field of the stationary magnet creates a magnetic torque on the magnet that is free to rotate. This magnetic torque tends to align a magnet's poles with the magnetic field lines. A compass, therefore, will turn to align itself with earth's magnetic field. Magnetic torque is used to drive electric motors. In one simple motor design, a magnet is fixed to a freely rotating shaft and subjected to a magnetic field from an array of electromagnets. By continuously switching the electric current through each of the electromagnets, thereby flipping the polarity of their magnetic fields, like poles are kept next to the rotor; the resultant torque is transferred to the shaft. See Rotating magnetic fields below. As is the case for the force between magnets, the magnetic pole model leads more readily to the correct equation. Here, two equal and opposite magnetic charges experiencing the same H also experience equal and opposite forces. Since these equal and opposite forces are in different locations, this produces a torque proportional to the distance (perpendicular to the force) between them. With the definition of m as the pole strength times the distance between the poles, this leads to = 0mHsin, where 0 is a constant called the magnetic constant and is the angle between H and m.

torque on a dipole: An H field (to right) causes equal but opposite forces on a N pole (+q) and a S pole (-q) creating a torque.

The Amperian loop model also predicts the same magnetic torque. Here, it is the B field interacting with the Amperian current loop through a Lorentz force described below. Again, the results are the same although the models are completely different. Mathematically, the torque on a small magnet is proportional both to the applied magnetic field and to the magnetic moment m of the magnet:

where represents the vector cross product. Note that this equation includes all of the qualitative information included above. There is no torque on a magnet if m is in the same direction as the magnetic field. (The cross product is zero for two vectors that are in the same direction.) Further, all other orientations feel a torque that twists them toward the direction of magnetic field.

Cross product: |a b| = a b sin.

Magnetic field

18

Magnetic field and electric currents


Currents of electric charges both generate a magnetic field and feel a force due to magnetic B-fields.

Magnetic field due to moving charges and electric currents


All moving charged particles produce magnetic fields. Moving point charges, such as electrons, produce complicated but well known magnetic fields that depend on the charge, velocity, and acceleration of the particles.[25] Magnetic field lines form in concentric circles around a cylindrical current-carrying conductor, such as a length of wire. The direction of such a magnetic field can be determined by using the "right hand grip rule" (see figure at right). The strength of the magnetic field decreases with distance from the wire. (For an infinite length wire the strength decreases inversely proportional to the distance.)

Right hand grip rule: a current flowing in the direction of the white arrow produces a magnetic field shown by the red arrows.

Bending a current-carrying wire into a loop concentrates the magnetic field inside the loop while weakening it outside. Bending a wire into multiple closely spaced loops to form a coil or "solenoid" enhances this effect. A device so formed around an iron core may act as an electromagnet, generating a strong, well-controlled magnetic field. An infinitely long cylindrical electromagnet has a uniform magnetic field inside, and no magnetic field outside. A finite length electromagnet produces a magnetic field that looks similar to that produced by a uniform permanent magnet, with its strength and polarity determined by the current flowing through the coil. The magnetic field generated by a steady current
Solenoid

(a constant flow of electric charges in which

charge is neither accumulating nor depleting at any point)[26] is described by the BiotSavart law:

where the integral sums over the wire length where vector d is the direction of the current, 0 is the magnetic constant, r is the distance between the location of d and the location at which the magnetic field is being calculated, and r is a unit vector in the direction of r. A slightly more general[27][28] way of relating the current to the B-field is through Ampre's law:

where the line integral is over any arbitrary loop and

enc

is the current enclosed by that loop. Ampre's law is

always valid for steady currents and can be used to calculate the B-field for certain highly symmetric situations such as an infinite wire or an infinite solenoid. In a modified form that accounts for time varying electric fields, Ampre's law is one of four Maxwell's equations that describe electricity and magnetism.

Magnetic field

19

Force on moving charges and current


Force on a charged particle A charged particle moving in a B-field experiences a sideways force that is proportional to the strength of the magnetic field, the component of the velocity that is perpendicular to the magnetic field and the charge of the particle. This force is known as the Lorentz force, and is given by

where F is the force, q is the electric charge of the particle, v is the instantaneous velocity of the particle, and B is the magnetic field (in teslas). The Lorentz force is always perpendicular to both the velocity of the particle and the magnetic field that created it. When a charged particle moves in a static magnetic field it will trace out a helical path in which the helix axis is parallel to the magnetic field and in which the speed of the particle will remain constant. Because the magnetic force is always perpendicular to the motion, the magnetic field Magnetic force on charged particles and currents, shown in 3d. The electric current shown here is conventional, the real current would be a charge of q can do no work on an isolated charge. It can flowing in exactly the opposite direction. Only one charge carrier is shown to only do work indirectly, via the electric field prevent cluttering the diagram. r1 is the position of entry into the field B, r2 is generated by a changing magnetic field. It is the exit. The vector l is the integral (sum) of all infinitesimal vectors dr from often claimed that the magnetic force can do r1 to r2. work to a non-elementary magnetic dipole, or to charged particles whose motion is constrained by other forces, but this is incorrect[29] because the work in those cases is performed by the electric forces of the charges deflected by the magnetic field. Force on current-carrying wire The force on a current carrying wire is similar to that of a moving charge as expected since a charge carrying wire is a collection of moving charges. A current carrying wire feels a force in the presence of a magnetic field. The Lorentz force on a macroscopic current is often referred to as the Laplace force. Consider a conductor of length l, cross section A, and charge q which is due to electric current i. If this conductor is placed in a magnetic field of induction B which makes an angle (theta) with the velocity of charges in the conductor, the force exerted on a single charge q is , so, for N charges where , the force exerted on the conductor is ,

Magnetic field , where . Direction of force The direction of force on a charge or a current can be determined by a mnemonic known as the right-hand rule. See the figure on the left. Using the right hand and pointing the thumb in the direction of the moving positive charge or positive current and the fingers in the direction of the magnetic field the resulting force on the charge points outwards from the palm. The force on a negatively charged particle is in the opposite direction. The right-hand rule: Pointing the thumb of the right hand in the If both the speed and the charge are reversed then the direction of the conventional current and the fingers in the direction direction of the force remains the same. For that reason of the B-field the force on the current points out of the palm. The force is reversed for a negative charge. a magnetic field measurement (by itself) cannot distinguish whether there is a positive charge moving to the right or a negative charge moving to the left. (Both of these cases produce the same current.) On the other hand, a magnetic field combined with an electric field can distinguish between these, see Hall effect below. An alternative mnemonic to the right hand rule is Fleming's left hand rule.

20

Relation between H and B


The formulas derived for the magnetic field above are correct when dealing with the entire current. A magnetic material placed inside a magnetic field, though, generates its own bound current which can be a challenge to calculate. (This bound current is due to the sum of atomic sized current loops and the spin of the subatomic particles such as electrons that make up the material.) The H-field as defined above helps factor out this bound current; but in order to see how, it helps to introduce the concept of magnetization first.

Magnetization
The magnetization vector field M represents how strongly a region of material is magnetized. It is defined as the net magnetic dipole moment per unit volume of that region. The magnetization of a uniform magnet, therefore, is a constant in the material equal to its magnetic moment, m, divided by its volume. Since the SI unit of magnetic moment is ampere-turn meter2, the SI unit of magnetization M is ampere-turn per meter which is identical to that of the H-field. The magnetization M field of a region points in the direction of the average magnetic dipole moment in that region. Magnetization field lines, therefore, begin near the magnetic south pole and ends near the magnetic north pole. (Magnetization does not exist outside of the magnet.) In the Amperian loop model, the magnetization is due to combining many tiny Amperian loops to form a resultant current called bound current. This bound current, then, is the source of the magnetic B field due to the magnet. (See Magnetic dipoles below and magnetic poles vs. atomic currents for more information.) Given the definition of the magnetic dipole, the magnetization field follows a similar law to that of Ampere's law:[30]

where the integral is a line integral over any closed loop and Ib is the 'bound current' enclosed by that closed loop. In the magnetic pole model, magnetization begins at and ends at magnetic poles. If a given region, therefore, has a net positive 'magnetic pole strength' (corresponding to a north pole) then it will have more magnetization field lines

Magnetic field entering it than leaving it. Mathematically this is equivalent to: , where the integral is a closed surface integral over the closed surface S and qM is the 'magnetic charge' (in units of magnetic flux) enclosed by S. (A closed surface completely surrounds a region with no holes to let any field lines escape.) The negative sign occurs because the magnetization field moves from south to north.

21

H-field and magnetic materials


The H-field is defined as: (definition of H in SI units) With this definition, Ampere's law becomes:

where If represents the 'free current' enclosed by the loop so that the line integral of H does not depend at all on the bound currents.[31] For the differential equivalent of this equation see Maxwell's equations. Ampere's law leads to the boundary condition where Kf is the surface free current density.[32] Similarly, a surface integral of H over any closed surface is independent of the free currents and picks out the 'magnetic charges' within that closed surface:

which does not depend on the free currents. The H-field, therefore, can be separated into two[33] independent parts:

where H0 is the applied magnetic field due only to the free currents and Hd is the demagnetizing field due only to the bound currents. The magnetic H-field, therefore, re-factors the bound current in terms of 'magnetic charges'. The H field lines loop only around 'free current' and, unlike the magnetic B field, begins and ends at near magnetic poles as well.

Magnetism
Most materials respond to an applied B-field by producing their own magnetization M and therefore their own B-field. Typically, the response is very weak and exists only when the magnetic field is applied. The term magnetism describes how materials respond on the microscopic level to an applied magnetic field and is used to categorize the magnetic phase of a material. Materials are divided into groups based upon their magnetic behavior: Diamagnetic materials[34] produce a magnetization that opposes the magnetic field. Paramagnetic materials[34] produce a magnetization in the same direction as the applied magnetic field. Ferromagnetic materials and the closely related ferrimagnetic materials and antiferromagnetic materials[35][36] can have a magnetization independent of an applied B-field with a complex relationship between the two fields. Superconductors (and ferromagnetic superconductors)[37][38] are materials that are characterized by perfect conductivity below a critical temperature and magnetic field. They also are highly magnetic and can be perfect diamagnets below a lower critical magnetic field. Superconductors often have a broad range of temperatures and magnetic fields (the so named mixed state) for which they exhibit a complex hysteretic dependence of M on B.

Magnetic field In the case of paramagnetism and diamagnetism, the magnetization M is often proportional to the applied magnetic field such that:

22

where is a material dependent parameter called the permeability. In some cases the permeability may be a second rank tensor so that H may not point in the same direction as B. These relations between B and H are examples of constitutive equations. However, superconductors and ferromagnets have a more complex B to H relation, see magnetic hysteresis.

Energy stored in magnetic fields


Energy is needed to generate a magnetic field both to work against the electric field that a changing magnetic field creates and to change the magnetization of any material within the magnetic field. For non-dispersive materials this same energy is released when the magnetic field is destroyed so that this energy can be modeled as being stored in the magnetic field. For linear, non-dispersive, materials (such that B = H where is frequency-independent), the energy density is:

If there are no magnetic materials around then can be replaced by 0. The above equation cannot be used for nonlinear materials, though; a more general expression given below must be used. In general, the incremental amount of work per unit volume W needed to cause a small change of magnetic field B is:

Once the relationship between H and B is known this equation is used to determine the work needed to reach a given magnetic state. For hysteretic materials such as ferromagnets and superconductors the work needed will also depend on how the magnetic field is created. For linear non-dispersive materials, though, the general equation leads directly to the simpler energy density equation given above.

Electromagnetism: the relationship between magnetic and electric fields


Faraday's Law: Electric force due to a changing B-field
A changing magnetic field, such as a magnet moving through a conducting coil, generates an electric field (and therefore tends to drive a current in the coil). This is known as Faraday's law and forms the basis of many electrical generators and electric motors. Mathematically, Faraday's law is:

where is the electromotive force (or EMF, the voltage generated around a closed loop) and m is the magnetic fluxthe product of the area times the magnetic field normal to that area. (This definition of magnetic flux is why B is often referred to as magnetic flux density.) The negative sign is necessary and represents the fact that any current generated by a changing magnetic field in a coil produces a magnetic field that opposes the change in the magnetic field that induced it. This phenomenon is known as Lenz's Law. This integral formulation of Faraday's law can be converted[39] into a differential form, which applies under slightly different conditions. This form is covered as one of Maxwell's equations below.

Magnetic field

23

Maxwell's correction to Ampre's Law: The magnetic field due to a changing electric field
Similar to the way that a changing magnetic field generates an electric field, a changing electric field generates a magnetic field. This fact is known as Maxwell's correction to Ampre's law. Maxwell's correction to Ampre's Law bootstrap together with Faraday's law of induction to form electromagnetic waves, such as light. Thus, a changing electric field generates a changing magnetic field which generates a changing electric field again. Maxwell's correction to Ampre law is applied as an additive term to Ampere's law given above. This additive term is proportional to the time rate of change of the electric flux and is similar to Faraday's law above but with a different and positive constant out front. (The electric flux through an area is proportional to the area times the perpendicular part of the electric field.) This full Ampre law including the correction term is known as the MaxwellAmpre equation. It is not commonly given in integral form because the effect is so small that it can typically be ignored in most cases where the integral form is used. The Maxwell term is critically important in the creation and propagation of electromagnetic waves. These, though, are usually described using the differential form of this equation given below.

Maxwell's equations
Like all vector fields, magnetic field has two important mathematical properties that relates it to its sources. (For the B-field the sources are currents and changing electric fields.) These two properties, along with the two corresponding properties of the electric field, make up Maxwell's Equations. Maxwell's Equations together with the Lorentz force law form a complete description of classical electrodynamics including both electricity and magnetism. The first property is the divergence of a vector field A, A which represents how A 'flows' outward from a given point. As discussed above, a B-field line never starts or ends at a point but instead forms a complete loop. This is mathematically equivalent to saying that the divergence of B is zero. (Such vector fields are called solenoidal vector fields.) This property is called Gauss's law for magnetism and is equivalent to the statement that there are no isolated magnetic poles or magnetic monopoles. The electric field on the other hand begins and ends at electric charges so that its divergence is non-zero and proportional to the charge density (See Gauss's law). The second mathematical property is called the curl, such that A represents how A curls or 'circulates' around a given point. The result of the curl is called a 'circulation source'. The equations for the curl of B and of E are called the AmpreMaxwell equation and Faraday's law respectively. They represent the differential forms of the integral equations given above. The complete set of Maxwell's equations then are:

where J = complete microscopic current density and is the charge density.

Magnetic field

24

Technically, B is a pseudovector (also called an axial vector) due to being defined by a vector cross product. (See diagram to right.) As discussed above, materials respond to an applied electric E field and an applied magnetic B field by producing their own internal 'bound' charge and current distributions that contribute to E and B but are difficult to calculate. To circumvent this problem, H and D fields are used to re-factor Maxwell's equations in terms of the free current density Jf and free charge density f:

Magnetic field, like all pseudovectors, changes sign when reflected in a mirror: When a current carrying loop (black) is reflected in a mirror (dotted line), its magnetic field (blue) is reflected and reversed.

These equations are not any more general than the original equations (if the 'bound' charges and currents in the material are known). They also need to be supplemented by the relationship between B and H as well as that between E and D. On the other hand, for simple relationships between these quantities this form of Maxwell's equations can circumvent the need to calculate the bound charges and currents.

Electric and magnetic fields: different aspects of the same phenomenon


According to the special theory of relativity, the partition of the electromagnetic force into separate electric and magnetic components is not fundamental, but varies with the observational frame of reference: An electric force perceived by one observer may be perceived by another (in a different frame of reference) as a magnetic force, or a mixture of electric and magnetic forces. Formally, special relativity combines the electric and magnetic fields into a rank-2 tensor, called the electromagnetic tensor. Changing reference frames mixes these components. This is analogous to the way that special relativity mixes space and time into spacetime, and mass, momentum and energy into four-momentum.

Magnetic vector potential


In advanced topics such as quantum mechanics and relativity it is often easier to work with a potential formulation of electrodynamics rather than in terms of the electric and magnetic fields. In this representation, the vector potential A, and the scalar potential , are defined such that:

The vector potential A may be interpreted as a generalized potential momentum per unit charge[40] just as is interpreted as a generalized potential energy per unit charge. Maxwell's equations when expressed in terms of the potentials can be cast into a form that agrees with special relativity with little effort.[41] In relativity A together with forms the four-potential analogous to the four-momentum which combines the momentum and energy of a particle. Using the four potential instead of the electromagnetic tensor has the advantage of being much simpler; further it can be easily modified to work with quantum mechanics.

Magnetic field

25

Quantum electrodynamics
In modern physics, the electromagnetic field is understood to be not a classical field, but rather a quantum field; it is represented not as a vector of three numbers at each point, but as a vector of three quantum operators at each point. The most accurate modern description of the electromagnetic interaction (and much else) is Quantum electrodynamics (QED),[42] which is incorporated into a more complete theory known as the Standard Model of particle physics. In QED, the magnitude of the electromagnetic interactions between charged particles (and their antiparticles) is computed using perturbation theory; these rather complex formulas have a remarkable pictorial representation as Feynman diagrams in which virtual photons are exchanged. Predictions of QED agree with experiments to an extremely high degree of accuracy: currently about 1012 (and limited by experimental errors); for details see precision tests of QED. This makes QED one of the most accurate physical theories constructed thus far. All equations in this article are in the classical approximation, which is less accurate than the quantum description mentioned here. However, under most everyday circumstances, the difference between the two theories is negligible.

Important uses and examples of magnetic field


Earth's magnetic field
The Earth's magnetic field is thought to be produced by convection currents in the outer liquid of Earth's core. The Dynamo theory proposes that these movements produce electric currents which, in turn, produce the magnetic field.[43] The presence of this field causes a compass, placed anywhere within it, to rotate so that the "north pole" of the magnet in the compass points roughly north, toward Earth's north magnetic pole. This is the traditional definition of the "north pole" of a magnet, although other equivalent definitions are also possible. One confusion that arises from this definition is that, if Earth itself is considered as a magnet, the south pole of that magnet would be the one nearer the north magnetic pole, and vice-versa[44] (opposite poles attract, so the north pole of the compass magnet is attracted to the south pole of Earth's interior magnet).

A sketch of Earth's magnetic field representing the source of the field as a magnet. The geographic north pole of Earth is near the top of the diagram, the south pole near the bottom. The south pole of that magnet is deep in Earth's interior below Earth's North Magnetic Pole.

The north magnetic pole is so-named not because of the polarity of the field there but because of its geographical location. The north and south poles of a permanent magnet are so-called because they are "north-seeking" and "south-seeking", respectively.[45] The figure to the right is a sketch of Earth's magnetic field represented by field lines. For most locations, the magnetic field has a significant up/down component in addition to the North/South component. (There is also an East/West component; Earth's magnetic poles do not coincide exactly with Earth's geological pole.) The magnetic field can be visualised as a bar magnet buried deep in Earth's interior. Earth's magnetic field is not constantthe strength of the field and the location of its poles vary. Moreover, the poles periodically reverse their orientation in a process called geomagnetic reversal. The most recent reversal occurred

Magnetic field 780,000 years ago.

26

Rotating magnetic fields


The rotating magnetic field is a key principle in the operation of alternating-current motors. A permanent magnet in such a field rotates so as to maintain its alignment with the external field. This effect was conceptualized by Nikola Tesla, and later utilized in his, and others', early AC (alternating-current) electric motors. A rotating magnetic field can be constructed using two orthogonal coils with 90 degrees phase difference in their AC currents. However, in practice such a system would be supplied through a three-wire arrangement with unequal currents. This inequality would cause serious problems in standardization of the conductor size and so, in order to overcome it, three-phase systems are used where the three currents are equal in magnitude and have 120 degrees phase difference. Three similar coils having mutual geometrical angles of 120 degrees create the rotating magnetic field in this case. The ability of the three-phase system to create a rotating field, utilized in electric motors, is one of the main reasons why three-phase systems dominate the world's electrical power supply systems. Because magnets degrade with time, synchronous motors use DC voltage fed rotor windings which allows the excitation of the machine to be controlled and induction motors use short-circuited rotors (instead of a magnet) following the rotating magnetic field of a multicoiled stator. The short-circuited turns of the rotor develop eddy currents in the rotating field of the stator, and these currents in turn move the rotor by the Lorentz force. In 1882, Nikola Tesla identified the concept of the rotating magnetic field. In 1885, Galileo Ferraris independently researched the concept. In 1888, Tesla gained U.S. Patent 381968 [46] for his work. Also in 1888, Ferraris published his research in a paper to the Royal Academy of Sciences in Turin.

Hall effect
The charge carriers of a current carrying conductor placed in a transverse magnetic field experience a sideways Lorentz force; this results in a charge separation in a direction perpendicular to the current and to the magnetic field. The resultant voltage in that direction is proportional to the applied magnetic field. This is known as the Hall effect. The Hall effect is often used to measure the magnitude of a magnetic field. It is used as well to find the sign of the dominant charge carriers in materials such as semiconductors (negative electrons or positive holes).

Magnetic circuits
An important use of H is in magnetic circuits where inside a linear material B = H. Here, is the magnetic permeability of the material. This result is similar in form to Ohm's law J = E, where J is the current density, is the conductance and E is the electric field. Extending this analogy, the counterpart to the macroscopic Ohm's law ( I = V R ) is:

where the circuit, and

is the magnetic flux in the circuit, is the reluctance of the circuit. Here the reluctance

is the magnetomotive force applied to is a quantity similar in nature to

resistance for the flux. Using this analogy it is straightforward to calculate the magnetic flux of complicated magnetic field geometries, by using all the available techniques of circuit theory.

Magnetic field

27

Magnetic field shape descriptions


An azimuthal magnetic field is one that runs east-west. A meridional magnetic field is one that runs north-south. In the solar dynamo model of the Sun, differential rotation of the solar plasma causes the meridional magnetic field to stretch into an azimuthal magnetic field, a process called the omega-effect. The reverse process is called the alpha-effect.[47] A dipole magnetic field is one seen around a bar magnet or around a charged elementary particle with nonzero spin. A quadrupole magnetic field is one seen, for example, between the poles of four bar magnets. The field strength grows linearly with the radial distance from its longitudinal axis. A solenoidal magnetic field is similar to a dipole magnetic field, except that a solid bar magnet is replaced by a hollow electromagnetic coil magnet.
Schematic quadrupole magnet ("four-pole") magnetic field. There are four steel pole tips, two opposing magnetic north poles and two opposing magnetic south poles.

A toroidal magnetic field occurs in a doughnut-shaped coil, the electric current spiraling around the tube-like surface, and is found, for example, in a tokamak. A poloidal magnetic field is generated by a current flowing in a ring, and is found, for example, in a tokamak. A radial magnetic field is one in which the field lines are directed from the center outwards, similar to the spokes in a bicycle wheel. An example can be found in a loudspeaker transducers (driver).[48] A helical magnetic field is corkscrew-shaped, and sometimes seen in space plasmas such as the Orion Molecular Cloud.[49]

Magnetic dipoles
The magnetic field of a magnetic dipole is depicted on the right. From outside, the ideal magnetic dipole is identical to that of an ideal electric dipole of the same strength. Unlike the electric dipole, a magnetic dipole is properly modeled as a current loop having a current I and an area a. Such a current loop has a magnetic moment of:

where the direction of m is perpendicular to the area of the loop and depends on the direction of the current using the right-hand rule. An ideal magnetic dipole is modeled as a real magnetic dipole whose area a has been reduced to zero and its current I increased to infinity such Magnetic field lines around a magnetostatic that the product m=Ia is finite. In this model it is easy to see the dipole pointing to the right. connection between angular momentum and magnetic moment which is the basis of the Einstein-de Haas effect "rotation by magnetization" and its inverse, the Barnett effect or "magnetization by rotation".[50] Rotating the loop faster (in the same direction) increases the current and therefore the magnetic moment, for example. It is sometimes useful to model the magnetic dipole similar to the electric dipole with two equal but opposite magnetic charges (one south the other north) separated by distance d. This model produces an H-field not a B-field. Such a model is deficient, though, both in that there are no magnetic charges and in that it obscures the link between electricity and magnetism. Further, as discussed above it fails to explain the inherent connection between angular momentum and magnetism.

Magnetic field

28

Magnetic monopole (hypothetical)


A magnetic monopole is a hypothetical particle (or class of particles) that has, as its name suggests, only one magnetic pole (either a north pole or a south pole). In other words, it would possess a "magnetic charge" analogous to an electric charge. Magnetic field lines would start or end on magnetic monopoles, so if they exist, they would give exceptions to the rule that magnetic field lines neither start nor end. Modern interest in this concept stems from particle theories, notably Grand Unified Theories and superstring theories, that predict either the existence, or the possibility, of magnetic monopoles. These theories and others have inspired extensive efforts to search for monopoles. Despite these efforts, no magnetic monopole has been observed to date.[51] In recent research, materials known as spin ices can simulate monopoles, but do not contain actual monopoles.

Notes
[1] Technically, a magnetic field is a pseudo vector; pseudo-vectors, which also include torque and rotational velocity, are similar to vectors except that they remain unchanged when the coordinates are inverted. [2] His Epistola Petri Peregrini de Maricourt ad Sygerum de Foucaucourt Militem de Magnete, which is often shortened to Epistola de magnete, is dated 1269 C.E. [3] Whittaker, E. T. (1951). A History of the Theories of Aether and Electricity (http:/ / www. archive. org/ details/ historyoftheorie00whitrich). Dover Publications. p.34. ISBN0486261263. . [4] in his "A Treatise of Artificial Magnets; in which is shown an easy and expeditious method of making them superior to the best natural ones" [5] Whittaker, E. T. (1951). A History of the Theories of Aether and Electricity (http:/ / www. archive. org/ details/ historyoftheorie00whitrich). Dover Publications. p.56. ISBN0486261263. . [6] Whittaker, E. T. (1951). A History of the Theories of Aether and Electricity (http:/ / www. archive. org/ details/ historyoftheorie00whitrich). Dover Publications. p.59. ISBN0486261263. . [7] Whittaker, E. T. (1951). A History of the Theories of Aether and Electricity (http:/ / www. archive. org/ details/ historyoftheorie00whitrich). Dover Publications. p.64. ISBN0486261263. . [8] Whittaker, E. T. (1951). A History of the Theories of Aether and Electricity (http:/ / www. archive. org/ details/ historyoftheorie00whitrich). Dover Publications. p.88. ISBN0486261263. . [9] It is a remarkable fact that from the 'outside' the field of a dipole of magnetic charge has the exact same form as that of an elementary current loop called a magnetic dipole. It is therefore only for the physics of magnetism 'inside' of magnetic material that the two models differ. [10] Electromagnetics, by Rothwell and Cloud, p23 (http:/ / books. google. com/ books?id=jCqv1UygjA4C& pg=PA23) [11] R.P. Feynman, R.B. Leighton, M. Sands (1963). The Feynman Lectures on Physics, volume 2. [12] Edward Purcell, in Electricity and Magnetism, McGraw-Hill, 1963, writes, Even some modern writers who treat B as the primary field feel obliged to call it the magnetic induction because the name magnetic field was historically preempted by H. This seems clumsy and pedantic. If you go into the laboratory and ask a physicist what causes the pion trajectories in his bubble chamber to curve, he'll probably answer "magnetic field", not "magnetic induction." You will seldom hear a geophysicist refer to the Earth's magnetic induction, or an astrophysicist talk about the magnetic induction of the galaxy. We propose to keep on calling B the magnetic field. As for H, although other names have been invented for it, we shall call it "the field H" or even "the magnetic field H." In a similar vein, M Gerloch (1983). Magnetism and Ligand-field Analysis (http:/ / books. google. com/ ?id=Ovo8AAAAIAAJ& pg=PA110). Cambridge University Press. p.110. ISBN0521249392. . says: "So we may think of both B and H as magnetic fields, but drop the word 'magnetic' from H so as to maintain the distinction ... As Purcell points out, 'it is only the names that give trouble, not the symbols'." [13] This can be seen from the magnetic part of the Lorentz force law Fmag = (qvB). [14] Magnetic Field Strength Converter (http:/ / www. unitconversion. org/ unit_converter/ magnetic-field-strength. html), UnitConversion.org. [15] "Gravity Probe B Executive Summary" (http:/ / www. nasa. gov/ pdf/ 168808main_gp-b_pfar_cvr-pref-execsum. pdf). pp.10, 21. . [16] "With record magnetic fields to the 21st Century" (http:/ / ieeexplore. ieee. org/ xpl/ freeabs_all. jsp?arnumber=823621). IEEE Xplore. . [17] Kouveliotou, C.; Duncan, R. C.; Thompson, C. (February 2003). " Magnetars (http:/ / solomon. as. utexas. edu/ ~duncan/ sciam. pdf)". Scientific American; Page 36. [18] The use of iron filings to display a field presents something of an exception to this picture; the filings alter the magnetic field so that it is much larger along the "lines" of iron, due to the large permeability of iron relative to air. [19] Here 'small' means that the observer is sufficiently far away that it can be treated as being infinitesimally small. 'Larger' magnets need to include more complicated terms in the expression and depend on the entire geometry of the magnet not just m. [20] Magnetic field lines may also wrap around and around without closing but also without ending. These more complicated non-closing non-ending magnetic field lines are moot, though, since the magnetic field of objects that produce them are calculated by adding the magnetic fields of 'elementary parts' having magnetic field lines that do form closed curves or extend to infinity. [21] To see that this must be true imagine placing a compass inside a magnet. There, the north pole of the compass points toward the north pole of the magnet since magnets stacked on each other point in the same direction.

Magnetic field
[22] As discussed above, magnetic field lines are primarily a conceptual tool used to represent the mathematics behind magnetic fields. The total 'number' of field lines is dependent on how the field lines are drawn. In practice, integral equations such as the one that follows in the main text are used instead. [23] Either B or H may be used for the magnetic field outside of the magnet. [24] See Eq. 11.42 in E. Richard Cohen, David R. Lide, George L. Trigg (2003). AIP physics desk reference (http:/ / books. google. com/ ?id=JStYf6WlXpgC& pg=PA381) (3 ed.). Birkhuser. p.381. ISBN0387989730. . [25] Griffiths, David J. (1999). Introduction to Electrodynamics (3rd ed.). Prentice Hall. p.438. ISBN0-13-805326-X. OCLC40251748. [26] In practice, the BiotSavart law and other laws of magnetostatics are often used even when the currents are changing in time as long as it is not changing too quickly. It is often used, for instance, for standard household currents which oscillate sixty times per second. [27] Griffiths, David J. (1999). Introduction to Electrodynamics (3rd ed.). Prentice Hall. pp.222225. ISBN0-13-805326-X. OCLC40251748. [28] The BiotSavart law contains the additional restriction (boundary condition) that the B-field must go to zero fast enough at infinity. It also depends on the divergence of B being zero, which is always valid. (There are no magnetic charges.) [29] Deissler, R.J. (2008). "Dipole in a magnetic field, work, and quantum spin" (http:/ / academic. csuohio. edu/ deissler/ PhysRevE_77_036609. pdf). Physical Review E 77 (3, pt 2): 036609. Bibcode2008PhRvE..77c6609D. doi:10.1103/PhysRevE.77.036609. PMID18517545. . [30] Griffiths, David J. (1999). Introduction to Electrodynamics (3rd ed.). Prentice Hall. pp.2668. ISBN0-13-805326-X. OCLC40251748. [31] John Clarke Slater, Nathaniel Herman Frank (1969). Electromagnetism (http:/ / books. google. com/ ?id=GYsphnFwUuUC& pg=PA69) (first published in 1947 ed.). Courier Dover Publications. p.69. ISBN0486622630. . [32] David Griffiths. Introduction to Electrodynamics (3rd 1999 ed.). p.332. [33] A third term is needed for changing electric fields and polarization currents; this displacement current term is covered in Maxwell's equations below. [34] RJD Tilley (2004). Understanding Solids (http:/ / books. google. com/ ?id=ZVgOLCXNoMoC& pg=PA368). Wiley. p.368. ISBN0470852755. . [35] Sshin Chikazumi, Chad D. Graham (1997). Physics of ferromagnetism (http:/ / books. google. com/ ?id=AZVfuxXF2GsC& printsec=frontcover) (2 ed.). Oxford University Press. p.118. ISBN0198517769. . [36] Amikam Aharoni (2000). Introduction to the theory of ferromagnetism (http:/ / books. google. com/ ?id=9RvNuIDh0qMC& pg=PA27) (2 ed.). Oxford University Press. p.27. ISBN0198508085. . [37] M Brian Maple et al. (2008). "Unconventional superconductivity in novel materials" (http:/ / books. google. com/ ?id=PguAgEQTiQwC& pg=PA640). In K. H. Bennemann, John B. Ketterson. Superconductivity. Springer. p.640. ISBN3540732527. . [38] Naoum Karchev (2003). "Itinerant ferromagnetism and superconductivity" (http:/ / books. google. com/ ?id=3AFo_yxBkD0C& pg=PA169). In Paul S. Lewis, D. Di (CON) Castro. Superconductivity research at the leading edge. Nova Publishers. p.169. ISBN1590338618. . [39] A complete expression for Faraday's law of induction in terms of the electric E and magnetic fields can be written as: where (t) is the moving closed path bounding the moving surface (t), and dA is an element of surface area of (t). The first integral calculates the work done moving a charge a distance d based upon the Lorentz force law. In the case where the bounding surface is stationary, the KelvinStokes theorem can be used to show this equation is equivalent to the MaxwellFaraday equation. [40] E. J. Konopinski (1978). "What the electromagnetic vector potential describes". Am. J. Phys. 46 (5): 499502. Bibcode1978AmJPh..46..499K. doi:10.1119/1.11298. [41] Griffiths, David J. (1999). Introduction to Electrodynamics (3rd ed.). Prentice Hall. p.422. ISBN0-13-805326-X. OCLC40251748. [42] For a good qualitative introduction see: Feynman, Richard (2006). QED: the strange theory of light and matter. Princeton University Press. ISBN0-691-12575-9. [43] Herbert, Yahreas (June 1954). "What makes the earth Wobble" (http:/ / books. google. com/ ?id=NiEDAAAAMBAJ& pg=PA96& dq=What+ makes+ the+ earth+ wobble& q=What makes the earth wobble). Popular Science (New York: Godfrey Hammond): 266. . [44] College Physics, Volume 10, by Serway, Vuille, and Faughn, page 628 weblink (http:/ / books. google. com/ books?id=CX0u0mIOZ44C& pg=PT660). "the geographic North Pole of Earth corresponds to a magnetic south pole, and the geographic South Pole of Earth corresponds to a magnetic north pole". [45] Kurtus, Ron (2004). "Magnets" (http:/ / www. school-for-champions. com/ science/ magnets. htm). School for champions: Physics topics. . Retrieved 17 July 2010. [46] http:/ / www. google. com/ patents?vid=381968 [47] The Solar Dynamo (http:/ / www. cora. nwra. com/ ~werne/ eos/ text/ dynamo. html), retrieved September 15, 2007. [48] I. S. Falconer and M. I. Large (edited by I. M. Sefton), " Magnetism: Fields and Forces (http:/ / www. physics. usyd. edu. au/ super/ life_sciences/ electricity. html)" Lecture E6, The University of Sydney, retrieved 3 October 2008 [49] Robert Sanders, " Astronomers find magnetic Slinky in Orion (http:/ / berkeley. edu/ news/ media/ releases/ 2006/ 01/ 12_helical. shtml)", 12 January 2006 at UC Berkeley. Retrieved 3 October 2008 [50] (See magnetic moment for further information.)

29

B. D. Cullity, C. D. Graham (2008). Introduction to Magnetic Materials (http:/ / books. google. com/ ?id=ixAe4qIGEmwC&pg=PA103) (2 ed.). Wiley-IEEE. p.103. ISBN0471477419. .

Magnetic field
[51] Two experiments produced candidate events that were initially interpreted as monopoles, but these are now regarded to be inconclusive. For details and references, see magnetic monopole.

30

References Further reading


Durney, Carl H. and Johnson, Curtis C. (1969). Introduction to modern electromagnetics. McGraw-Hill. ISBN0-07-018388-0. Furlani, Edward P. (2001). Permanent Magnet and Electromechanical Devices: Materials, Analysis and Applications. Academic Press Series in Electromagnetism. ISBN0-12-269951-3. OCLC162129430. Jiles, David (1994). Introduction to Electronic Properties of Materials (1st ed ed.). Springer. ISBN0-412-49580-5. Kraftmakher, Yaakov (2001). "Two experiments with rotating magnetic field" (http://www.iop.org/EJ/ abstract/0143-0807/22/5/302). Eur. J. Phys. 22: 477482. Melle, Sonia; Rubio, Miguel A.; Fuller, Gerald G. (2000). "Structure and dynamics of magnetorheological fluids in rotating magnetic fields" (http://prola.aps.org/abstract/PRE/v61/i4/p4111_1). Phys. Rev. E 61: 41114117. Bibcode2000PhRvE..61.4111M. doi:10.1103/PhysRevE.61.4111. Rao, Nannapaneni N. (1994). Elements of engineering electromagnetics (4th ed.). Prentice Hall. ISBN0-13-948746-8. OCLC221993786. Mielnik, Bogdan (1989). "An electron trapped in a rotating magnetic field" (http://scitation.aip.org/getabs/ servlet/GetabsServlet?prog=normal&id=JMAPAQ000030000002000537000001&idtype=cvips&gifs=yes). Journal of Mathematical Physics 30 (2): 537549. Bibcode1989JMP....30..537M. doi:10.1063/1.528419. Thalmann, Julia K. (2010). Evolution of Coronal Magnetic Fields. uni-edition. ISBN978-3-942171-41-0. Tipler, Paul (2004). Physics for Scientists and Engineers: Electricity, Magnetism, Light, and Elementary Modern Physics (5th ed.). W. H. Freeman. ISBN0-7167-0810-8. OCLC51095685.

External links
Information
Crowell, B., " Electromagnetism (http:/ / www. lightandmatter. com/ html_books/ 0sn/ ch11/ ch11. html)". Nave, R., " Magnetic Field (http:/ / hyperphysics. phy-astr. gsu. edu/ hbase/ magnetic/ magfie. html)". HyperPhysics. "Magnetism", The Magnetic Field (http:/ / theory. uwinnipeg. ca/ physics/ mag/ node2. html#SECTION00110000000000000000). theory.uwinnipeg.ca. Hoadley, Rick, " What do magnetic fields look like (http:/ / my. execpc. com/ ~rhoadley/ magfield. htm)?" 17 July 2005.

Rotating magnetic fields


" Rotating magnetic fields (http:/ / www. tpub. com/ neets/ book5/ 18a. htm)". Integrated Publishing. "Introduction to Generators and Motors", rotating magnetic field (http:/ / www. tpub. com/ content/ neets/ 14177/ css/ 14177_87. htm). Integrated Publishing. " Induction Motor Rotating Fields (http:/ / web. archive. org/ web/ 20050929102550/ http:/ / www. egr. msu. edu/ ~jurkovi4/ Experiment4. pdf)".

Field density
Oppelt, Arnulf (2 November 2006). "magnetic field strength" (http:/ / searchsmb. techtarget. com/ sDefinition/ 0,290660,sid44_gci763586,00. html). Retrieved 04 June 2007. "magnetic field strength converter" (http:/ / www. unitconversion. org/ unit_converter/ magnetic-field-strength. html). Retrieved 04 June 2007.

Diagrams
"AC Motor Theory" Figure 2 Rotating Magnetic Field (http:/ / www. tpub. com/ content/ doe/ h1011v4/ css/ h1011v4_23. htm). Integrated Publishing. "Magnetic Fields" Arc & Mitre Magnetic Field Diagrams (http:/ / www. first4magnets. com/ ekmps/ shops/ trainer27/ resources/ Other/ magnetic-fields. pdf). Magnet Expert Ltd.

Magnetization

31

Magnetization
In classical electromagnetism, magnetization [1] or magnetic polarization is the vector field that expresses the density of permanent or induced magnetic dipole moments in a magnetic material. The origin of the magnetic moments responsible for magnetization can be either microscopic electric currents resulting from the motion of electrons in atoms, or the spin of the electrons or the nuclei. Net magnetization results from the response of a material to an external magnetic field, together with any unbalanced magnetic dipole moments that may be inherent in the material itself; for example, in ferromagnets. Magnetization is not always homogeneous within a body, but rather varies between different points. Magnetization also describes how a material responds to an applied magnetic field as well as the way the material changes the magnetic field, and can be used to calculate the forces that result from those interactions. It can be compared to electric polarization, which is the measure of the corresponding response of a material to an electric field in electrostatics. Physicists and engineers define magnetization as the quantity of magnetic moment per unit volume. It is represented by a vector M.

Definition
Magnetization can be defined according to the following equation:

Here, M represents magnetization; m is the vector that defines the magnetic moment; V represents volume; and N is the number of magnetic moments in the sample. The quantity N/V is usually written as n, the number density of magnetic moments. The M-field is measured in amperes per meter (A/m) in SI units.[2]

Magnetization in Maxwell's equations


The behavior of magnetic fields (B, H), electric fields (E, D), charge density (), and current density (J) is described by Maxwell's equations. The role of the magnetization is described below.

Relations between B, H, and M


The magnetization defines the auxiliary magnetic field H as (SI units) (Gaussian units) which is convenient for various calculations. The vacuum permeability 0 is, by definition, 4107 Vs/(Am). A relation between M and H exists in many materials. In diamagnets and paramagnets, the relation is usually linear:

where m is called the volume magnetic susceptibility. In ferromagnets there is no one-to-one correspondence between M and H because of hysteresis.

Magnetization

32

Magnetization current
The magnetization M makes a contribution to the current density J, known as the magnetization current or bound current:

so that the total current density that enters Maxwell's equations is given by

where Jf is the electric current density of free charges (also called the free current), the second term is the contribution from the magnetization, and the last term is related to the electric polarization P.

Magnetostatics
In the absence of free electric currents and time-dependent effects, Maxwell's equations describing the magnetic quantities reduce to

These equations can be easily solved in analogy with electrostatic problems where

In this sense

plays the role of a "magnetic charge density" analogous to the electric charge density

(see

also demagnetizing field). Magnetization is volume density of magnetic moment. That is: if a certain volume has magnetization volume element has a magnetic moment of then the

Magnetization dynamics
Main article: Magnetization dynamics The time-dependent behavior of magnetization becomes important when considering nanoscale and nanosecond timescale magnetization. Rather than simply aligning with an applied field, the individual magnetic moments in a material begin to precess around the applied field and come into alignment through relaxation as energy is transferred into the lattice.

Demagnetization
In addition to magnetization, there is also demagnetization. Demagnetization is the process by which the magnetic field of an object is reduced or eliminated.[3] The process of demagnetizing can be accomplished in many ways. One technique used for demagnetization is to heat the object above its Curie Temperature. The reason for this is that when a magnetic material is heated to its Curie Temperature, the material's magnetivity is eliminated. One other way of achieving demagnetization is to use an electric coil. If the object is retracted out of a coil with aternating current running through it, the object's dipoles will become randomized and the object will be demagnetized.[4]

Magnetization

33

Applications of Demagnetization
One application of demagnetization is to eliminate unwanted magnetic fields. The reason for doing this is that magnetic fields can have unwanted effects on different devices. In particular magnetic fields can affect electronic devices such as cell phones or computers. If such a device is going to be coming into contact with other possibly magnetic objects, the magnetic fields might need to be reduced in order to protect the electronic device. Therefore demagnetization is sometimes used to keep magnetic fields from damaging electrical devices.[5]

Sources
[1] American spelling. The British spelling is magnetisation. [2] "Units for Magnetic Properties" (http:/ / www. magneticmicrosphere. com/ resources/ Units_for_Magnetic_Properties. pdf). Lake Shore Cryotronics, Inc.. . Retrieved 2009-10-24. [3] "Magnetic Component Engineering" (http:/ / www. mceproducts. com/ knowledge-base/ article/ article-dtl. asp?id=90). Magnetic Component Engineering. . Retrieved April 18, 2011. [4] "Demagnetization" (http:/ / www. ndt-ed. org/ EducationResources/ CommunityCollege/ MagParticle/ Physics/ Demagnetization. htm). Introduction to Magnetic Particle Inspection. NDT Resource Center. . Retrieved April 18, 2011. [5] "Demagnetization" (http:/ / www. ndt-ed. org/ EducationResources/ CommunityCollege/ MagParticle/ Physics/ Demagnetization. htm). Introduction to Magnetic Particle Inspection. NDT Resource Center. . Retrieved April 18, 2011.

Magnetic moment
The magnetic moment of a magnet is a quantity that determines the force that the magnet can exert on electric currents and the torque that a magnetic field will exert on it. A loop of electric current, a bar magnet, an electron, a molecule, and a planet all have magnetic moments. Both the magnetic moment and magnetic field may be considered to be vectors having a magnitude and direction. The direction of the magnetic moment points from the south to north pole of a magnet. The magnetic field produced by a magnet is proportional to its magnetic moment as well. More precisely, the term magnetic moment normally refers to a system's magnetic dipole moment, which produces the first term in the multipole expansion of a general magnetic field. The dipole component of an object's magnetic field is symmetric about the direction of its magnetic dipole moment, and decreases as the inverse cube of the distance from the object.

Magnetic moment

34

Two definitions of moment


The preferred definition of a magnetic moment has changed over time. Before the 1930's, textbooks defined the moment using magnetic poles. Since then, most have defined it in terms of Amprian currents.[1]

Magnetic pole definition


The sources of magnetic moments in materials can be represented by poles in analogy to electrostatics. Consider a bar magnet which has magnetic poles of equal magnitude but opposite polarity. Each pole is the source of magnetic force which weakens with distance. Since magnetic poles always come in pairs, their forces partially cancel each other because while one pole pulls, the other repels. This cancellation is greatest when the poles are close to each other i.e. when the bar magnet is short. The magnetic force produced by a bar magnet, at a given point in space, therefore depends on two factors: on both the strength p of its poles, and on the vector separating them. The moment is defined as[1]

It points in the direction from South to North pole. The analogy with electric dipoles should not be taken too far because magnetic dipoles are associated with angular momentum (see Magnetic moment and angular momentum). Nevertheless, magnetic poles are very useful for magnetostatic calculations, particularly in applications to ferromagnets.[1] Practitioners using the magnetic pole approach generally represent the magnetic field by the irrotational field H, in analogy to the electric field E.

An electrostatic analogue for a magnetic moment: two opposing charges separated by a finite distance.

Current loop definition


Suppose a planar closed loop carries an electric current I and has vector area S (x, y, and z coordinates of this vector are the areas of projections of the loop onto the yz, zx, and xy planes). Its magnetic moment m, vector, is defined as:

Moment of a planar current having magnitude I and enclosing an area S.

By convention, the direction of the vector area is given by the right hand grip rule (curling the fingers of one's right hand in the direction of the current around the loop, when the palm of the hand is "touching" the loop's outer edge, and the straight thumb indicates the direction of the vector area and thus of the magnetic moment). [2]

If the loop is not planar, the moment is given as

where is the vector cross product. In the most general case of an arbitrary current distribution in space, the magnetic moment of such a distribution can be found from the following equation:

where r is the position vector pointing from the origin to the location of the volume element, and J is the current density vector at that location.

Magnetic moment The above equation can be used for calculating a magnetic moment of any assembly of moving charges, such as a spinning charged solid, by substituting

35

where is the electric charge density at a given point and v is the instantaneous linear velocity of that point. For example, the magnetic moment produced by an electric charge moving along a circular path is , where r is the position of the charge q relative to the center of the circle and v is the instantaneous velocity of the charge. Practitioners using the current loop model generally represent the magnetic field by the solenoidal field B, analogous to the electrostatic field D. Magnetic moment of a solenoid A generalization of the above current loop is a multi-turn coil, or solenoid. Its moment is the vector sum of the moments of individual turns. If the solenoid has N identical turns (single-layer winding),

Units
The unit for magnetic moment is not a base unit in the International System of Units (SI) and it can be represented in more than one way. For example, in the current loop definition, the area is measured in square meters and I is measured in amperes, so the magnetic moment is 3-D image of a solenoid. measured in amperesquare meters (A m2). In the equation for torque on a moment, the torque is measured in joules and the magnetic field in tesla, so the moment is measured in Joules per Tesla (JT1). These two representations are equivalent: 1 Am2 = 1 JT1. In the CGS system, there are several different sets of electromagnetism units, of which the main ones are ESU, Gaussian, and EMU. Among these, there are two alternative (non-equivalent) units of magnetic dipole moment in CGS: (ESU CGS) 1 stat Acm2 = 3.335640951014 (Am2 or JT1) and (more frequently used) (EMU CGS and Gaussian-CGS) 1 ergG = 1 abAcm2 = 103 (m2A or J/T). The ratio of these two non-equivalent CGS units (EMU/ESU) is equal exactly to the speed of light in free space, expressed in cms1. All formulas in this article are correct in SI units, but in other unit systems, the formulas may need to be changed. For example, in SI units, a loop of current with current I and area A has magnetic moment IA (see below), but in Gaussian units the magnetic moment is IA/c.

Magnetic moment

36

Effects of an external magnetic field on a magnetic moment


Force on a moment
A magnetic moment in an externally-produced magnetic field has a potential energy U:

In a case when the external magnetic field is non-uniform, there will be a force, proportional to the magnetic field gradient, acting on the magnetic moment itself. There has been some discussion on how to calculate the force acting on a magnetic dipole. There are two expressions for the force acting on a magnetic dipole, depending on whether the model used for the dipole is a current loop or two monopoles (analogous to the electric dipole).[3] The force obtained in the case of a current loop model is

In the case of a pair of monopoles being used (i.e. electric dipole model)

and one can be put in terms of the other via the relation

In all these expressions m is the dipole and B is the magnetic field at its position. Note that if there are no currents or time-varying electrical fields B = 0 and the two expressions agree. An electron, nucleus, or atom placed in a uniform magnetic field will precess with a frequency known as the Larmor frequency. See Resonance.

Torque on a moment
The magnetic moment can also be defined as a vector relating the aligning torque on the object from an externally applied magnetic field to the field vector itself. The relationship is given by [4]

where is the torque acting on the dipole and B is the external magnetic field.

Magnetic moment and angular momentum


The magnetic moment has a close connection with angular momentum called the gyromagnetic effect. This effect is expressed on a macroscopic scale in the Einstein-de Haas effect, or "rotation by magnetization," and its inverse, the Barnett effect, or "magnetization by rotation."[4] In particular, when a magnetic moment is subject to a torque in a magnetic field that tends to align it with the applied magnetic field, the moment precesses (rotates about the axis of the applied field). This is a consequence of the angular momentum associated with the moment. Viewing a magnetic dipole as a rotating charged sphere brings out the close connection between magnetic moment and angular momentum. Both the magnetic moment and the angular momentum increase with the rate of rotation of the sphere. The ratio of the two is called the gyromagnetic ratio, usually denoted by the symbol .[5] [6] For a spinning charged solid with a uniform charge density to mass density ratio, the gyromagnetic ratio is equal to half the charge-to-mass ratio. This implies that a more massive assembly of charges spinning with the same angular momentum will have a proportionately weaker magnetic moment, compared to its lighter counterpart. Even though atomic particles cannot be accurately described as spinning charge distributions of uniform charge-to-mass ratio, this general trend can be observed in the atomic world, where the intrinsic angular momentum (spin) of each type of particle is a constant: a small half-integer times the reduced Planck constant . This is the basis for defining the magnetic moment units of Bohr magneton (assuming charge-to-mass ratio of the electron) and nuclear magneton (assuming charge-to-mass ratio of the proton).

Magnetic moment

37

Magnetic dipoles
A magnetic dipole is the limit of either a current loop or a pair of poles as the dimensions of the source are reduced to zero while keeping the moment constant. As long as these limits only applies to fields far from the sources, they are equivalent. However, the two models give different predictions for the internal field (see below).

External magnetic field produced by a magnetic dipole moment


Any system possessing a net magnetic dipole moment m will produce a dipolar magnetic field (described below) in the space surrounding the system. While the net magnetic field produced by the system can also have higher-order multipole components, those will drop off with distance more rapidly, so that only the dipolar component will dominate the magnetic field of the system at distances far away from it. The vector potential of magnetic field produced by magnetic moment m is

and magnetic flux density is


Magnetic field lines around a magnetostatic dipole the magnetic dipole itself is in the center and is seen from the side.

Alternatively one can obtain the scalar potential first from the magnetic pole perspective,

and hence magnetic field strength is

The magnetic field of an ideal magnetic dipole is depicted on the left.

Magnetic moment

38

Internal magnetic field of a dipole


The two models for a dipole (current loop and magnetic poles) give the same predictions for the magnetic field far from the source. However, inside the source region they give different predictions. The magnetic field between poles (see figure for Magnetic pole definition) is in the opposite direction to the magnetic moment (which points from the negative charge to the positive charge), while inside a current loop it is in the same direction (see the figure to the right). Clearly, the limits of these fields must also be different as the sources shrink to zero size. This distinction only matters if the dipole limit is used to calculate fields inside a magnetic material. If a magnetic dipole is formed by making a current loop smaller and smaller, but keeping the product of current and area constant, the limiting field is

The magnetic field of a current loop.

Unlike the expressions in the previous section, this limit is correct for the internal field of the dipole. If a magnetic dipole is formed by taking a "north pole" and a "south pole", bringing them closer and closer together but keeping the product of magnetic pole-charge and distance constant, the limiting field is

These fields are related by is the magnetization.

, where

Forces between two magnetic dipoles


If B in the previous equations is replaced with the expression of the field of a magnetic dipole under the approximation for distances bigger than the characteristic length of the dipole.[7] Namely,

where the variables r and are Frames of reference for calculating the forces between two dipoles measured in a frame of reference with origin in m1 and oriented in such a way that m1 lies in the x-axis. This frame is called Local coordinates and is shown in the Figure on the right. The final formulas are shown next. They are expressed in the global coordinate system,

Magnetic moment

39

Using vector notation, the above equations can be written as [8]

where r is the distance-vector from dipole moment m1 to dipole moment m2, with r = ||r||, and where F is the force acting on m2. The force acting on m1 is in opposite direction. The torque is straightforward to obtain from the formula

which gives (in global coordinates)

or

in local coordinates.

Examples of magnetic moments


Two kinds of magnetic sources
Fundamentally, contributions to any system's magnetic moment may come from sources of two kinds: (1) motion of electric charges, such as electric currents, and (2) the intrinsic magnetism of elementary particles, such as the electron. Contributions due to the sources of the first kind can be calculated from knowing the distribution of all the electric currents (or, alternatively, of all the electric charges and their velocities) inside the system, by using the formulas below. On the other hand, the magnitude of each elementary particle's intrinsic magnetic moment is a fixed number, often measured experimentally to a great precision. For example, any electron's magnetic moment is measured to be 9.2847641024 J/T.[9] The direction of the magnetic moment of any elementary particle is entirely determined by the direction of its spin (the minus in front of the value above indicates that any electron's magnetic moment is antiparallel to its spin). The net magnetic moment of any system is a vector sum of contributions from one or both types of sources. For example, the magnetic moment of an atom of hydrogen-1 (the lightest hydrogen isotope, consisting of a proton and an electron) is a vector sum of the following contributions: 1. the intrinsic moment of the electron, 2. the orbital motion of the electron around the proton, 3. the intrinsic moment of the proton. Similarly, the magnetic moment of a bar magnet is the sum of the intrinsic and orbital magnetic moments of the unpaired electrons of the magnet's material.

Magnetic moment

40

Magnetic moment of an atom


For an atom, individual electron spins are added to get a total spin, and individual orbital angular momenta are added to get a total orbital angular momentum. These two then are added using angular momentum coupling to get a total angular momentum. The magnitude of the atomic dipole moment is then[10]

where J is the total angular momentum quantum number, gJ is the Land g-factor, and B is the Bohr magneton. The component of this magnetic moment along the direction of the magnetic field is then[11]

where m is called the magnetic quantum number or the equatorial quantum number, which can take on any of 2J+1 values: [12] . The negative sign occurs because electrons have negative charge. Due to the angular momentum, the dynamics of a magnetic dipole in a magnetic field differs from that of an electric dipole in an electric field. The field does exert a torque on the magnetic dipole tending to align it with the field. However, torque is proportional to rate of change of angular momentum, so precession occurs: the direction of spin changes. This behavior is described by the Landau-Lifshitz-Gilbert equation:[13][14]

where is gyromagnetic ratio, m is magnetic moment, is damping coefficient and Heff is effective magnetic field (the external field plus any self-field). The first term describes precession of the moment about the effective field, while the second is a damping term related to dissipation of energy caused by interaction with the surroundings.

Magnetic moment of an electron


Electrons and many elementary particles also have intrinsic magnetic moments, an explanation of which requires a quantum mechanical treatment and relates to the intrinsic angular momentum of the particles as discussed in the article electron magnetic dipole moment. It is these intrinsic magnetic moments that give rise to the macroscopic effects of magnetism, and other phenomena, such as electron paramagnetic resonance. The magnetic moment of the electron is

where B is the Bohr magneton, S is electron spin, and the electron g-factor gS is 2 according to Dirac's theory, but due to quantum electrodynamic effects it is slightly larger in reality: 2.002 319 304 36. Again it is important to notice that m is a negative constant multiplied by the spin, so the magnetic moment of the electron is antiparallel to the spin. This can be understood with the following classical picture: if we imagine that the spin angular momentum is created by the electron mass spinning around some axis, the electric current that this rotation creates circulates in the opposite direction, because of the negative charge of the electron; such current loops produce a magnetic moment which is antiparallel to the spin. Hence, for a positron (the anti-particle of the electron) the magnetic moment is parallel to its spin.

Magnetic moment

41

Magnetic moment of a nucleus


The nuclear system is a complex physical system consisting of nucleons, i.e., protons and neutrons. The quantum mechanical properties of the nucleons include the spin among others. Since the electromagnetic moments of the nucleus depend on the spin of the individual nucleons, one can look at these properties with measurements of nuclear moments, and more specifically the nuclear magnetic dipole moment. Most common nuclei exist in their ground state, although nuclei of some isotopes have long-lived excited states. Each energy state of a nucleus of a given isotope is characterized by a well-defined magnetic dipole moment, the magnitude of which is a fixed number, often measured experimentally to a great precision. This number is very sensitive to the individual contributions from nucleons, and a measurement or prediction of its value can reveal important information about the content of the nuclear wave function. There are several theoretical models that predict the value of the magnetic dipole moment and a number of experimental techniques aiming to carry out measurements in nuclei along the nuclear chart.

Magnetic moment of a molecule


Any molecule has a well-defined magnitude of magnetic moment, which may depend on the molecule's energy state. Typically, the overall magnetic moment of a molecule is a combination of the following contributions, in the order of their typical strength: magnetic moments due to its unpaired electron spins (paramagnetic contribution), if any orbital motion of its electrons, which in the ground state is often proportional to the external magnetic field (diamagnetic contribution) the combined magnetic moment of its nuclear spins, which depends on the nuclear spin configuration. Examples of molecular magnetism Oxygen molecule, O2, exhibits strong paramagnetism, due to unpaired spins of its outermost two electrons. Carbon dioxide molecule, CO2, mostly exhibits diamagnetism, a much weaker magnetic moment of the electron orbitals that is proportional to the external magnetic field. In the rare instance when a magnetic isotope, such as 13 C or 17O, is present, it will contribute its nuclear magnetism to the molecule's magnetic moment. Hydrogen molecule, H2, in a weak (or zero) magnetic field exhibits nuclear magnetism, and can be in a para- or an ortho- nuclear spin configuration.

Elementary particles
In atomic and nuclear physics, the symbol represents the magnitude of the magnetic moment, often measured in Bohr magnetons or nuclear magnetons, associated with the intrinsic spin of the particle and/or with the orbital motion of the particle in a system. Values of the intrinsic magnetic moments of some particles are given in the table below:

Magnetic moment

42

Intrinsic magnetic moments and spins of some elementary particles [15]


Particle electron proton neutron muon deuteron triton helion Magnetic dipole moment in SI units (1027 JT1) Spin quantum number (dimensionless) -9284.764 14.106067 -9.66236 -44.904478 4.3307346 15.046094 -10.746174 1/2 1/2 1/2 1/2 1 1/2 1/2 0

alpha particle 0

For relation between the notions of magnetic moment and magnetization see magnetization.

References and notes


[1] Brown, Jr., William Fuller. Magnetostatic Principles in Ferromagnetism. North-Holland. [2] Feynman, Richard P.; Leighton, Robert B.; Sands, Matthew (2006). The Feynman Lectures on Physics. 2. ISBN0-8053-9045-6. [3] Boyer, Timothy H. (1988). "The Force on a Magnetic Dipole". American Journal of Physics 56 (8): 688692. Bibcode1988AmJPh..56..688B. doi:10.1119/1.15501. [4] B. D. Cullity, C. D. Graham (2008). Introduction to Magnetic Materials (http:/ / books. google. com/ ?id=ixAe4qIGEmwC& pg=PA103) (2 ed.). Wiley-IEEE Press. p.103. ISBN0471477419. . [5] Uwe Krey, Anthony Owen (2007). Basic Theoretical Physics (http:/ / books. google. com/ ?id=xZ_QelBmkxYC& pg=PA151). Springer. pp.151152. ISBN3540368043. . [6] Richard B. Buxton (2002). Introduction to functional magnetic resonance imaging (http:/ / books. google. com/ ?id=6XVu0NKzgekC& pg=PA136). Cambridge University Press. p.136. ISBN0521581133. . [7] Schill, R. A. (2003). "General relation for the vector magnetic field of a circular current loop: A closer look". IEEE Transactions on Magnetics 39 (2): 961967. Bibcode2003ITM....39..961S. doi:10.1109/TMAG.2003.808597. [8] Furlani, Edward P. (2001). Permanent Magnet and Electromechanical Devices: Materials, Analysis, and Applications (http:/ / books. google. com/ ?id=irsdLnC5SrsC& dq=permanent+ magnet+ and+ electromechanical+ devices& printsec=frontcover& q=3. 130). Academic Press. p.140. ISBN0122699513. . [9] "CODATA value: electron magnetic moment" (http:/ / physics. nist. gov/ cgi-bin/ cuu/ Value?muem). . [10] RJD Tilley (2004). Understanding Solids (http:/ / books. google. com/ ?id=ZVgOLCXNoMoC& pg=PA368). John Wiley and Sons. p.368. ISBN0470852755. . [11] Paul Allen Tipler, Ralph A. Llewellyn (2002). Modern Physics (http:/ / books. google. com/ ?id=tpU18JqcSNkC& pg=PA310) (4 ed.). Macmillan. p.310. ISBN0716743450. . [12] JA Crowther (2007). Ions, Electrons and Ionizing Radiations (http:/ / books. google. com/ ?id=H_sft9-zm5AC& pg=PA277) (reprinted Cambridge (1934) 6 ed.). Rene Press. p.277. ISBN1406720399. . [13] Stuart Alan Rice (2004). Advances in chemical physics (http:/ / books. google. com/ ?id=wK3Vhq-VnBQC& pg=PA208). Wiley. pp.208 ff. ISBN0471445282. . [14] Marcus Steiner (2004). Micromagnetism and Electrical Resistance of Ferromagnetic Electrodes for Spin Injection Devices (http:/ / books. google. com/ ?id=tnX1edkCB-wC& pg=PA6). Cuvillier Verlag. p.6. ISBN3865371760. . [15] See NIST's Fundamental Physical Constants website http:/ / physics. nist. gov/ cgi-bin/ cuu/ Results?search_for=+ magnetic+ moment

Demagnetizing field

43

Demagnetizing field
The demagnetizing field, also called the stray field, is the magnetic field (H-field)[1] generated by the magnetization in a magnet. The total magnetic field in a region containing magnets is the sum of the demagnetizing fields of the magnets and the magnetic field due to any free currents or displacement currents. The term demagnetizing field reflects its tendency to act on the magnetization so as to reduce the total magnetic moment. It gives rise to shape anisotropy in ferromagnets with a single magnetic domain and to magnetic domains in larger ferromagnets. The demagnetizing field of an arbitrary shaped object is very difficult to calculate even for the simple case of uniform magnetization. For the special case of ellipsoids (which includes spheres) the demagnetization field is linearly related to the magnetization by a geometry dependent constant called the demagnetizing factor. Since the magnetization of a sample at a given location depends on the total magnetic field at that point, the demagnetization factor must be used in order to accurately determine how a magnetic material responds to a magnetic field. (See magnetic hysteresis.)

Magnetostatic principles
Maxwell's equations
In general the demagnetizing field is a function of position H(r). It is derived from the magnetostatic equations for a body with no electric currents.[2] These are Ampre's law
[3] (1)

and Gauss's law


[4] (2)

The magnetic field and flux density are related by[5][6]


[7] (3)

The magnetic potential


The general solution of the first equation can be expressed as the gradient of a scalar potential U(r):
[5][6] (4)

Inside the magnetic body, the potential Uin is determined by substituting (3) and (4) in (2):
[8] (5)

Outside the body, where the magnetization is zero,


(6)

At the surface of the magnet, there are two continuity requirements:[5] The component of H parallel to the surface must be continuous (no jump in value at the surface). The component of B perpendicular to the surface must be continuous.

Demagnetizing field This leads to the following boundary conditions at the surface of the magnet:

44

(7)

Here n is the surface normal and

is the derivative with respect to distance from the surface.[9]

The outer potential Uout must also be regular at infinity: both |r U| and |r2 U| must be bounded as r goes to infinity. This ensures that the magnetic energy is finite.[10] Sufficiently far away, the magnetic field looks like the field of a magnetic dipole with the same moment as the finite body.

Uniqueness of the demagnetizing field


Any two potentials that satisfy equations (5), (6) and (7), along with regularity at infinity, are identical. The demagnetizing field Hd is the gradient of this potential (equation 4).

Energy
The energy of the demagnetizing field is completely determined by an integral over the volume V of the magnet:
(7)

Suppose there are two magnets with magnetizations M1 and M2. The energy of the first magnet in the demagnetizing field Hd(2) of the second is
(8)

The reciprocity theorem states that[9]


(9)

Magnetic charge and the pole-avoidance principle


Formally, the solution of the equations for the potential is
(10)

where r is the variable to be integrated over the volume of the body in the first integral and the surface in the second, and is the gradient with respect to this variable.[9] Qualitatively, the negative of the divergence of the magnetization M (called a volume pole) is analogous to a bulk bound electric charge in the body while n M (called a surface pole) is analogous to a bound surface electric charge. Although the magnetic charges do not exist, it can be useful to think of them in this way. In particular, the arrangement of magnetization that reduces the magnetic energy can often be understood in terms of the pole-avoidance principle, which states that the magnetization tries to reduce the poles as much as possible.[9]

Demagnetizing field

45

Effect on magnetization
Single domain
One way to remove the magnetic poles inside a ferromagnet is to make the magnetization uniform. This occurs in single-domain ferromagnets. This still leaves the surface poles, so division into domains reduces the poles further. However, very small ferromagnets are kept uniformly magnetized by the exchange interaction. The concentration of poles depends on the direction of magnetization (see the figure). If the magnetization is along the longest axis, the poles are spread across a smaller surface, so the energy is lower. This is a form of magnetic anisotropy called shape anisotropy.

Illustration of the magnetic charges at the surface of a single-domain ferromagnet. The arrows indicate the direction of magnetization. The thickness of the colored region indicates the surface charge density.

Multiple domains
If the ferromagnet is large enough, its magnetization can divide into domains. It is then possible to have the magnetization parallel to the surface. Within each domain the magnetization is uniform, so there are no volume poles, but there are surface poles at the interfaces (domain walls) between domains. However, these poles vanish if the magnetic moments on each side of the domain wall meet the wall at the same angle (so that the components n M are the same but opposite in sign). Domains configured this way are called closure domains.
Illustration of a magnet with four magnetic closure domains. The magnetic charges contributed by each domain are pictured at one domain wall. The charges balance, so the total charge is zero.

Demagnetizing factor

An arbitrary shaped magnetic object has a total magnetic field that varies with location inside the object and can be quite difficult to calculate. This makes it very difficult to determine the magnetic properties of a material such as, for instance, how the magnetization of a material varies with the magnetic field. For a uniformly magnetized sphere in a uniform magnetic field H0 the internal magnetic field H is uniform:
(11)

where M0 is the magnetization of the sphere and is called the demagnetizing factor and equals 4/3 for a sphere.[5][6][11] This equation can be generalized to include ellipsoids having principal axes in x,y, and z directions such that each component has a relationship of the form:[6]
(12)

Demagnetizing field Other important examples are an infinite plate (an ellipsoid with two of its axes going to infinity) which has = 4 in a direction normal to the plate and zero otherwise and an infinite cylinder (an ellipsoid with one of its axes tending toward infinity with the other two being the same) which has = 2 along its axis. For tables or equations for the magnetizing factors of the general ellipsoid see Osborn 1945.

46

Notes
[1] In this article the term 'magnetic field' is used for the magnetic 'H field' while 'magnetic flux density' is used for the magnetic 'B-field'. [2] If there are electric currents in the system, they can be calculated separately and added to the solutions of these equations. [3] In words, the curl of the magnetic field is zero. [4] In words, the divergence of the magnetic flux density is zero. [5] Jackson 1975, chapter 5 [6] Nayfeh & Brussel 1985, chapter 9 [7] SI units are used in this article. [8] The symbol 2 is the Laplace operator. [9] Aharoni 1996, chapter 6 [10] Brown, Jr. 1962 [11] Griffiths 1999, chapter 6

References
Aharoni, Amikam (1996). Introduction to the Theory of Ferromagnetism (http://www.oup.com/us/catalog/ general/subject/Physics/ElectricityMagnetism/?view=usa&ci=9780198508090). Clarendon Press. ISBN0198517912. Brown, Jr., William Fuller (1962). Magnetostatic Principles in Ferromagnetism. Interscience. Griffiths, David J. (1999). Introduction to Electrodynamics (third ed.). Prentice Hall. ISBN9780138053260. Jackson, John David (1975). Classical Electrodynamics (Second ed.). John Wiley & Sons. ISBN0-471-43132-X. Nayfeh, Munir H.; Brussel, Morton K. (1985). Electricity and Magnetism. John Wiley & Sons. ISBN0-471-87681-X. Osborn, J. A. (1945). "Demagnetizing Factors of the General Ellipsoid" (http://www.cmap.polytechnique.fr/ ~alouges/coursm2/Osborn.pdf). Physical Review 67 (1112): 3517. Bibcode1945PhRv...67..351O. doi:10.1103/PhysRev.67.351.

Magnetic susceptibility

47

Magnetic susceptibility
In electromagnetism, the magnetic susceptibility (latin: susceptibilis receptiveness) is a dimensionless proportionality constant that indicates the degree of magnetization of a material in response to an applied magnetic field. A related term is magnetizability, the proportion between magnetic moment and magnetic flux density.[1] A closely related parameter is the permeability, which expresses the total magnetization of material and volume.

Definition of volume susceptibility


See also Relative permeability. The volume magnetic susceptibility, represented by the symbol (often simply , sometimes magnetic, to distinguish from the electric susceptibility), is defined by the relationship

where, in SI units, M is the magnetization of the material (the magnetic dipole moment per unit volume), measured in amperes per meter, and H is the magnetic field strength, also measured in amperes per meter. The magnetic induction B is related to H by the relationship

where 0 is the magnetic constant (see table of physical constants), and material. Thus the volume magnetic susceptibility following formula: .

is the relative permeability of the are related by the

and the magnetic permeability

Sometimes an auxiliary quantity, called intensity of magnetization (also referred to as magnetic polarisation J) and measured in teslas, is defined as . This allows an alternative description of all magnetization phenomena in terms of the quantities I and B, as opposed to the commonly used M and H.

[2]

Conversion between SI and CGS units


Note that these definitions are according to SI conventions. However, many tables of magnetic susceptibility give CGS values (more specifically emu-cgs, short for electromagnetic units, or Gaussian-cgs; both are the same in this context) that rely on a different definition of the permeability of free space:[3]

The dimensionless CGS value of volume susceptibility is multiplied by 4 to give the dimensionless SI volume susceptibility value:[3] For example, the CGS volume magnetic susceptibility of water at 20C is 7.19107 which is 9.04106 using the SI convention.

Magnetic susceptibility

48

Mass susceptibility and molar susceptibility


There are two other measures of susceptibility, the mass magnetic susceptibility (mass or g, sometimes m), measured in m3kg1 in SI or in cm3g1 in CGS and the molar magnetic susceptibility (mol) measured in m3mol1 (SI) or cm3mol1 (CGS) that are defined below, where is the density in kgm3 (SI) or gcm3 (CGS) and M is molar mass in kgmol1 (SI) or gmol1 (CGS).

Sign of susceptibility: diamagnetics and other types of magnetism


If is positive, the material can be paramagnetic . In this case, the magnetic field in the material is strengthened by the induced magnetization. Alternatively, if is negative, the material is diamagnetic. As a result, the magnetic field in the material is weakened by the induced magnetization. Generally, non-magnetic materials are said para- or diamagnetic because they do not possess permanent magnetization without external magnetic field. Ferromagnetic, ferrimagnetic, or antiferromagnetic materials, which have positive susceptibility, possess permanent magnetization even without external magnetic field.

Experimental methods to determine susceptibility


Volume magnetic susceptibility is measured by the force change felt upon the application of a magnetic field gradient.[4] Early measurements were made using the Gouy balance where a sample is hung between the poles of an electromagnet. The change in weight when the electromagnet is turned on is proportional to the susceptibility. Today, high-end measurement systems use a superconductive magnet. An alternative is to measure the force change on a strong compact magnet upon insertion of the sample. This system, widely used today, is called the Evans balance.[5] For liquid samples, the susceptibility can be measured from the dependence of the NMR frequency of the sample on its shape or orientation.[6][7][8][9][10]

Tensor susceptibility
The magnetic susceptibility of most crystals is not a scalar. Magnetic response M is dependent upon the orientation of the sample and can occur in directions other than that of the applied field H. In these cases, volume susceptibility is defined as a tensor

where i and j refer to the directions (e.g., x and y in Cartesian coordinates) of the applied field and magnetization, respectively. The tensor is thus rank 2 (second order), dimension (3,3) describing the component of magnetization in the i-th direction from the external field applied in the j-th direction.

Differential susceptibility
In ferromagnetic crystals, the relationship between M and H is not linear. To accommodate this, a more general definition of differential susceptibility is used

where

is a tensor derived from partial derivatives of components of M with respect to components of H. When

the coercivity of the material parallel to an applied field is the smaller of the two, the differential susceptibility is a function of the applied field and self interactions, such as the magnetic anisotropy. When the material is not saturated, the effect will be nonlinear and dependent upon the domain wall configuration of the material.

Magnetic susceptibility

49

Susceptibility in the frequency domain


When the magnetic susceptibility is measured in response to an AC magnetic field (i.e. a magnetic field that varies sinusoidally), this is called AC susceptibility. AC susceptibility (and the closely related "AC permeability") are complex quantities, and various phenomena (such as resonances) can be seen in AC susceptibility that cannot in constant-field (DC) susceptibility. In particular, when an ac-field is applied perpendicular to the detection direction (called the "transverse susceptibility" regardless of the frequency), the effect has a peak at the ferromagnetic resonance frequency of the material with a given static applied field. Currently, this effect is called the microwave permeability or network ferromagnetic resonance in the literature. These results are sensitive to the domain wall configuration of the material and eddy currents. In terms of ferromagnetic resonance, the effect of an ac-field applied along the direction of the magnetization is called parallel pumping. For a tutorial with more information on AC susceptibility measurements, see here (external link) [11].

Examples
Magnetic susceptibility of some materials
Material Temperature Pressure (molar susc.) (mass susc.) (volume susc.) M (molar mass) (density)

Units

(C)

(atm)

SI (m mol1)
3

CGS (cm3mol1)

SI (m kg1)
3

CGS (cm3g1)

SI

CGS (emu)

(103 (103 kg/m3) kg/mol) or (g/cm3) or (g/mol) 0 0.9982

vacuum water [12] bismuth [13] Diamond [14] Graphite [15] (to c-axis) Graphite [15]

Any 20

0 1

1.6311010 1.298105 9.051109 7.203107 9.035106 7.190107 18.015

20

3.55109

2.82104

1.70108

1.35106

1.66104

1.32105

208.98

9.78

R.T.

7.41011

5.9106

6.2109

4.9107

2.2105

1.7106

12.01

3.513

R.T.

7.51011

6.0106

6.3109

5.0107

1.4105

1.1106

12.01

2.267

R.T.

3.2109

2.6104

2.7107

2.2105

6.1104

4.9105

12.01

2.267

Graphite [15]

-173

4.4109

3.5104

3.6107

2.9105

8.3104

6.6105

12.01

2.267

He Xe O2

[16] [16] [16]

20 20 20

1 1 0.209

2.381011 5.711010 4.3108

1.89106 4.54105 3.42103

5.93109 4.35109 1.34106

4.72107 3.46107 1.07104

9.851010 7.841011 4.0026 2.37108 3.73107 1.89109 2.97108 131.29 31.99

0.000166 0.00546 0.000278

Magnetic susceptibility
[16]

50
0.781 1 28.01 26.98 107.87 0.000910 2.70

N2 Al Ag

20

1.561010 2.21010

1.24105 1.7105

5.56109 7.9109

4.43107 6.3107

5.06109 2.2105 2.31105

4.031010 1.75106 1.84106

[17]

961

Sources of confusion in published data


There are tables of magnetic susceptibility values published on-line that seem to have been uploaded from a substandard source,[18] which itself has probably borrowed heavily from the CRC Handbook of Chemistry and Physics. Some of the data (e.g. for Al, Bi, and diamond) are apparently in cgs Molar Susceptibility units, whereas that for water is in Mass Susceptibility units (see discussion above). The susceptibility table in the CRC Handbook is known to suffer from similar errors, and even to contain sign errors. Effort should be made to trace the data in such tables to the original sources, and to double-check the proper usage of units.

References and notes


[1] "magnetizability, " (http:/ / goldbook. iupac. org/ search. py?search_text=magnetizability). IUPAC Compendium of Chemical TerminologyThe Gold Book (2nd ed.). International Union of Pure and Applied Chemistry. 1997. . [2] Richard A. Clarke. "Magnetic properties of materials" (http:/ / info. ee. surrey. ac. uk/ Workshop/ advice/ coils/ mu/ #itns). Info.ee.surrey.ac.uk. . Retrieved 2011-11-08. [3] Bennett, L. H.; Page, C. H.; and Swartzendruber, L. J. (1978). "Comments on units in magnetism". Journal of Research of the National Bureau of Standards (NIST, USA) 83 (1): 912. [4] L. N. Mulay (1972). A. Weissberger and B. W. Rossiter. ed. Techniques of Chemistry. 4. Wiley-Interscience: New York. p.431. [5] "Magnetic Susceptibility Balances" (http:/ / www. sherwood-scientific. com/ msb/ msbindex. html). Sherwood-scientific.com. . Retrieved 2011-11-08. [6] J. R. Zimmerman, and M. R. Foster (1957). "Standardization of NMR high resolution spectra". J. Phys. Chem. 61 (3): 282289. doi:10.1021/j150549a006. [7] Robert Engel, Donald Halpern, and Susan Bienenfeld (1973). "Determination of magnetic moments in solution by nuclear magnetic resonance spectrometry". Anal. Chem. 45 (2): 367369. doi:10.1021/ac60324a054. [8] P. W. Kuchel, B. E. Chapman, W. A. Bubb, P. E. Hansen, C. J. Durrant, and M. P. Hertzberg (2003). "Magnetic susceptibility: solutions, emulsions, and cells". Concepts Magn. Reson. A 18: 5671. doi:10.1002/cmr.a.10066. [9] K. Frei and H. J. Bernstein (1962). "Method for determining magnetic susceptibilities by NMR". J. Chem. Phys. 37 (8): 18911892. Bibcode1962JChPh..37.1891F. doi:10.1063/1.1733393. [10] R. E. Hoffman (2003). "Variations on the chemical shift of TMS". J. Magn. Reson. 163 (2): 325331. Bibcode2003JMagR.163..325H. doi:10.1016/S1090-7807(03)00142-3. PMID12914848. [11] http:/ / www. qdusa. com/ resources/ pdf/ 1078-201. pdf [12] G. P. Arrighini, M. Maestro, and R. Moccia (1968). "Magnetic Properties of Polyatomic Molecules: Magnetic Susceptibility of H2O, NH3, CH4, H2O2". J. Chem. Phys. 49 (2): 882889. Bibcode1968JChPh..49..882A. doi:10.1063/1.1670155. [13] S. Otake, M. Momiuchi and N. Matsuno (1980). "Temperature Dependence of the Magnetic Susceptibility of Bismuth". J. Phys. Soc. Jap. 49 (5): 18241828. Bibcode1980JPSJ...49.1824O. doi:10.1143/JPSJ.49.1824. The tensor needs to be averaged over all orientations: . [14] J. Heremans, C. H. Olk and D. T. Morelli (1994). "Magnetic Susceptibility of Carbon Structures". Phys. Rev. B 49 (21): 1512215125. Bibcode1994PhRvB..4915122H. doi:10.1103/PhysRevB.49.15122. [15] N. Ganguli and K.S. Krishnan (1941). "The Magnetic and Other Properties of the Free Electrons in Graphite". Proc. R. Soc. London 177 (969): 168182. Bibcode1941RSPSA.177..168G. doi:10.1098/rspa.1941.0002. [16] R. E. Glick (1961). "On the Diamagnetic Susceptibility of Gases". J. Phys. Chem. 65 (9): 15521555. doi:10.1021/j100905a020. [17] R. Dupree and C. J. Ford (1973). "Magnetic susceptibility of the noble metals around their melting points". Phys. Rev. B 8 (4): 17801782. Bibcode1973PhRvB...8.1780D. doi:10.1103/PhysRevB.8.1780. [18] "Magnetic Properties Susceptibilities Chart from" (http:/ / www. reade. com/ Particle_Briefings/ magnetic_susceptibilities. html). READE. 2006-01-11. . Retrieved 2011-11-08.

Permeability (electromagnetism)

51

Permeability (electromagnetism)
In electromagnetism, permeability is the measure of the ability of a material to support the formation of a magnetic field within itself. In other words, it is the degree of magnetization that a material obtains in response to an applied magnetic field. Magnetic permeability is typically represented by the Greek letter . The term was coined in September, 1885 by Oliver Heaviside. The reciprocal of magnetic permeability is magnetic reluctivity. We can simplify it by saying, the more conductive a material is to a magnetic field, the higher its permeability.

In SI units, permeability is measured in the henries per meter (Hm1), ferromagnets (f), paramagnets(p), free space(0) and diamagnets (d) or newtons per ampere squared (NA2). The permeability constant (0), also known as the magnetic constant or the permeability of free space, is a measure of the amount of resistance encountered when forming a magnetic field in a classical vacuum. The magnetic constant has the exact (defined)[1] value 0 = 4107 1.2566370614106Hm1 or NA2). A closely related property of materials is magnetic susceptibility, which is a measure of the magnetization of a material in addition to the magnetization of the space occupied by the material.

Simplified comparison of permeabilities for:

Explanation
In electromagnetism, the auxiliary magnetic field H represents how a magnetic field B influences the organization of magnetic dipoles in a given medium, including dipole migration and magnetic dipole reorientation. Its relation to permeability is

where the permeability is a scalar if the medium is isotropic or a second rank tensor for an anisotropic medium. In general, permeability is not a constant, as it can vary with the position in the medium, the frequency of the field applied, humidity, temperature, and other parameters. In a nonlinear medium, the permeability can depend on the strength of the magnetic field. Permeability as a function of frequency can take on real or complex values. In ferromagnetic materials, the relationship between B and H exhibits both non-linearity and hysteresis: B is not a single-valued function of H,[2] but depends also on the history of the material. For these materials it is sometimes useful to consider the incremental permeability defined as . This definition is useful in local linearizations of non-linear material behavior, for example in a Newton-Raphson iterative solution scheme that computes the changing saturation of a magnetic circuit. Permeability is the inductance per unit length. In SI units, permeability is measured in henries per metre (Hm1= J/(A2m)= N A2). The auxiliary magnetic field H has dimensions current per unit length and is measured in units of amperes per metre (A m1). The product H thus has dimensions inductance times current per unit area (HA/m2). But inductance is magnetic flux per unit current, so the product has dimensions magnetic flux per unit area. This is just the magnetic field B, which is measured in webers (volt-seconds) per square-metre (Vs/m2), or teslas (T). B is related to the Lorentz force on a moving charge q: . The charge q is given in coulombs (C), the velocity v in m/s, so that the force F is in newtons (N):

Permeability (electromagnetism)

52

H is related to the magnetic dipole density. A magnetic dipole is a closed circulation of electric current. The dipole moment has dimensions current times area, units ampere square-metre (Am2), and magnitude equal to the current around the loop times the area of the loop.[3] The H field at a distance from a dipole has magnitude proportional to the dipole moment divided by distance cubed,[4] which has dimensions current per unit length.

Relative permeability
Relative permeability, sometimes denoted by the symbol r, is the ratio of the permeability of a specific medium to the permeability of free space given by the magnetic constant . In terms of relative permeability, the magnetic susceptibility is: . m, a dimensionless quantity, is sometimes called volumetric or bulk susceptibility, to distinguish it from p (magnetic mass or specific susceptibility) and M (molar or molar mass susceptibility). :

Diamagnetism
Diamagnetism is the property of an object which causes it to create a magnetic field in opposition of an externally applied magnetic field, thus causing a repulsive effect. Specifically, an external magnetic field alters the orbital velocity of electrons around their nuclei, thus changing the magnetic dipole moment in the direction opposing the external field. Diamagnets are materials with a magnetic permeability less than (a relative permeability less than 1). Consequently, diamagnetism is a form of magnetism that a substance exhibits only in the presence of an externally applied magnetic field. It is generally a quite weak effect in most materials, although superconductors exhibit a strong effect.

Paramagnetism
Paramagnetism is a form of magnetism which occurs only in the presence of an externally applied magnetic field. Paramagnetic materials are attracted to magnetic fields, hence have a relative magnetic permeability greater than one (or, equivalently, a positive magnetic susceptibility). The magnetic moment induced by the applied field is linear in the field strength and rather weak. It typically requires a sensitive analytical balance to detect the effect. Unlike ferromagnets, paramagnets do not retain any magnetization in the absence of an externally applied magnetic field, because thermal motion causes the spins to become randomly oriented without it. Thus the total magnetization will drop to zero when the applied field is removed. Even in the presence of the field there is only a small induced magnetization because only a small fraction of the spins will be oriented by the field. This fraction is proportional to the field strength and this explains the linear dependency. The attraction experienced by ferromagnets is non-linear and much stronger, so that it is easily observed, for instance, in magnets on one's refrigerator.

Permeability (electromagnetism)

53

Values for some common materials


The following table should be used with caution as the permeability of ferromagnetic materials varies greatly with field strength. For example 4% Si steel has an initial relative permeability (at or near 0T) of 2,000 and a maximum of 35,000[5] and, indeed, the relative permeability of any material at a sufficiently high field strength tends to 1.

Magnetic susceptibility and permeability data for selected materials


Medium Susceptibility m Permeability [H/m] Relative Permeability /0 Magnetic field Frequency max. (volumetric SI) 1.25 10102 2.5102 1000000 80000 20000 50000 1.0102 5.0103 8000 4000 [6] at 0.5 T at 0.5 T at 0.002 T 100kHz 10kHz

Metglas Nanoperm Mu-metal Mu-metal Permalloy Electrical steel Ferrite (nickel zinc) Ferrite (manganese zinc) Steel Nickel Neodymium magnet Platinum Aluminum Wood Air Concrete Vacuum Hydrogen Teflon Sapphire Copper 2.1107 0 2.2109 [11] 2.22105 [11]

[7] [8] [9]

[8] [8]

at 0.002 T at 0.002 T 100kHz ~ 1MHz 100kHz ~ 1MHz at 0.002 T

2.0105 8.0104 16640 >8.0104 8.75104 1.25104 640 (or more) 100 100 [8] [8] 600

at 0.002 T

1.05 1.2569701106 1.2566650106

[10]

1.000265 1.000022 1.00000043 1.00000037 1 [13] [14] [11] [12]

1.2566371106 (0) 1.2566371106 1.2567106 [8]

1.0000000 1.0000 0.99999976 0.999994

1.2566368106

6.4106 1.2566290106 6[11] or 9.210 8.0106 1.66104 1 0 1.2566270106

Water Bismuth Superconductors

0.999992 0.999834 0

Permeability (electromagnetism) A good magnetic core material must have high permeability.[15] For passive magnetic levitation a relative permeability below 1 is needed (corresponding to a negative susceptibility). Permeability varies with magnetic field. Values shown above are approximate and valid only at the magnetic fields shown. Moreover, they are given for a zero frequency; in practice, the permeability is generally a function of the frequency. When frequency is considered the permeability can be complex, corresponding to the in phase and out of phase response. Note that the magnetic constant has an exact value in SI units (that

54

Magnetisation curve for ferromagnets (and ferrimagnets) and corresponding permeability

is, there is no uncertainty in its value), because the definition of the ampere fixes its value to 4107H/m exactly.

Complex permeability
A useful tool for dealing with high frequency magnetic effects is the complex permeability. While at low frequencies in a linear material the magnetic field and the auxiliary magnetic field are simply proportional to each other through some scalar permeability, at high frequencies these quantities will react to each other with some lag time.[16] These fields can be written as phasors, such that

where

is the phase delay of

from

. Understanding permeability as the ratio of the magnetic field to the

auxiliary magnetic field, the ratio of the phasors can be written and simplified as , so that the permeability becomes a complex number. By Euler's formula, the complex permeability can be translated from polar to rectangular form, . The ratio of the imaginary to the real part of the complex permeability is called the loss tangent, , which provides a measure of how much power is lost in a material versus how much is stored.

References
[1] "The NIST reference on fundamental physical constants" (http:/ / physics. nist. gov/ cuu/ Units/ ampere. html). Physics.nist.gov. . Retrieved 2011-11-08. [2] Jackson (1975), p. 190 [3] Jackson, John David (1975). Classical Electrodynamics (2nd ed. ed.). New York: Wiley. ISBN0-471-43132-X. p. 182 eqn. (5.57) [4] Jackson (1975) p. 182 eqn. (5.56) [5] G.W.C. Kaye & T.H. Laby, Table of Physical and Chemical Constants, 14th ed, Longman [6] ""Metglas Magnetic Alloy 2714A", ''Metglas''" (http:/ / www. metglas. com/ products/ page5_1_2_6. htm). Metglas.com. . Retrieved 2011-11-08. [7] ""Typical material properties of NANOPERM", ''Magnetec''" (http:/ / www. magnetec. de/ eng/ pdf/ werkstoffkennlinien_nano_e. pdf) (PDF). . Retrieved 2011-11-08. [8] ""Relative Permeability", ''Hyperphysics''" (http:/ / hyperphysics. phy-astr. gsu. edu/ hbase/ solids/ ferro. html). Hyperphysics.phy-astr.gsu.edu. . Retrieved 2011-11-08. [9] "Nickel Alloys-Stainless Steels, Nickel Copper Alloys, Nickel Chromium Alloys, Low Expansion Alloys" (http:/ / www. nickel-alloys. net/ nickelalloys. html). Nickel-alloys.net. . Retrieved 2011-11-08.

Permeability (electromagnetism)
[10] Juha Pyrhnen, Tapani Jokinen, Valria Hrabovcov (2009). Design of Rotating Electrical Machines (http:/ / books. google. com/ ?id=_y3LSh1XTJYC& pg=PT232). John Wiley and Sons. p.232. ISBN0-470-69516-1. . [11] Richard A. Clarke. "Clarke, R. ''Magnetic properties of materials'', surrey.ac.uk" (http:/ / www. ee. surrey. ac. uk/ Workshop/ advice/ coils/ mu/ ). Ee.surrey.ac.uk. . Retrieved 2011-11-08. [12] B. D. Cullity and C. D. Graham (2008), Introduction to Magnetic Materials, 2nd edition, 568 pp., p.16 [13] NDT.net. "Determination of dielectric properties of insitu concrete at radar frequencies" (http:/ / www. ndt. net/ article/ ndtce03/ papers/ v078/ v078. htm). Ndt.net. . Retrieved 2011-11-08. [14] exactly, by definition [15] Dixon, L H (2001). "Magnetics Design 2 - Magnetic Core Characteristics" (http:/ / www. ti. com/ lit/ ml/ slup124/ slup124. pdf). Texas Instruments. . [16] M. Getzlaff, Fundamentals of magnetism, Berlin: Springer-Verlag, 2008.

55

External links
Electromagnetism (http://www.lightandmatter.com/html_books/0sn/ch11/ch11.html) - a chapter from an online textbook Relative Permeability (http://hyperphysics.phy-astr.gsu.edu/hbase/solids/ferro.html) Soil Permeability Test (http://www.denichsoiltest.com) Magnetic Properties of Materials (http://www.ee.surrey.ac.uk/Workshop/advice/coils/mu/)

Force between magnets


Magnets exert forces and torques on each other due to the complex rules of electromagnetism. The forces of attraction field of magnets are due to microscopic currents of electrically charged electrons orbiting nuclei and the intrinsic magnetism of fundamental particles (such as electrons) that make up the material. Both of these are modeled quite well as tiny loops of current called magnetic dipoles that produce their own magnetic field and are affected by external magnetic fields. The most elementary force between magnets, therefore, is the magnetic dipoledipole interaction. If all of the magnetic dipoles that make up two magnets are known then the net force on both magnets can be determined by summing up all these interaction between the dipoles of the first magnet and that of the second. It is always more convenient to model the force between two magnets as being due to forces between magnetic poles having magnetic charges 'smeared' over them. Such a model fails to account for many important properties of magnetism such as the relationship between angular momentum and magnetic dipoles. Further, magnetic charge does not exist. This model works quite well, though, in predicting the forces between simple magnets where good models of how the 'magnetic charge' is distributed is available.

Force between magnets

56

Magnetic poles vs. atomic currents


Two models are used to calculate the magnetic fields of and the forces between magnets. The physically correct method is called the Ampre model while the easier model to use is often the Gilbert model. Ampre model: In the Ampre model, all magnetization is due to the effect of microscopic, or atomic, circular bound currents, also called Amprian currents throughout the material. The net effect of these microscopic bound currents is to make the magnet behave as if there is a macroscopic electric current flowing in loops in the magnet with the magnetic field normal to the loops. The Ampre model gives Field of a cylindrical bar magnet calculated with Ampre's the exact magnetic field both inside and outside the magnet. It model is usually difficult to calculate the Amprian currents on the surface of a magnet, though, whereas it is often easier to find the effective poles for the same magnet. Gilbert model: However, a version of the magnetic pole approach is used by professional magneticians to design permanent magnets. In this approach, the pole surfaces of a permanent magnet are imagined to be covered with so-called magnetic charge, north pole particles on the north pole and south pole particles' on the south pole, that are the source of the magnetic field lines. If the magnetic pole distribution is known, then outside the magnet the pole model gives the magnetic field exactly. In the interior of the magnet this model fails to give the correct field. This pole model is also called the Gilbert model of a magnetic dipole.[1] Griffiths suggests (p.258): "My advice is to use the Gilbert model, if you like, to get an intuitive 'feel' for a problem, but never rely on it for quantitative results."

Magnetic dipole moment


Far away from a magnet, the magnetic field created by that magnet is almost always described (to a good approximation) by a dipole field characterized by its total magnetic dipole moment, m. This is true regardless of the shape of the magnet, so long as the magnetic moment is non-zero. One characteristic of a dipole field is that the strength of the field falls off inversely with the cube of the distance from the magnet's center. The magnetic moment, therefore, of a magnet is a measure of its strength and orientation. A loop of electric current, a bar magnet, an electron, a molecule, and a planet all have magnetic moments. More precisely, the term magnetic moment normally refers to a system's magnetic dipole moment, which produces the first term in the multipole expansion[2] of a general magnetic field. Both the torque and force exerted on a magnet by an external magnetic field are proportional to that magnet's magnetic moment. The magnetic moment, like the magnetic field it produces, is a vector field; it has both a magnitude and direction. The direction of the magnetic moment points from the south to north pole of a magnet. For example the direction of the magnetic moment of a bar magnet, such as the one in a compass it the direction that the north poles points toward. In the physically correct Ampre model, magnetic dipole moments are due to infinitesimally small loops of current. For a sufficiently small loop of current, I, and area, A, the magnetic dipole moment is: , where the direction of m is normal to the area in a direction determined using the current and the right-hand rule. As such, the SI unit of magnetic dipole moment is amperemetre2. More precisely, to account for solenoids with many turns the unit of magnetic dipole moment is Ampere-turnmetre2.

Force between magnets In the Gilbert model, the magnetic dipole moment is due to two equal and opposite magnetic charges that are separated by a distance, d. In this model, m is similar to the electric dipole moment p due to electrical charges: , where qm is the 'magnetic charge'. The direction of the magnetic dipole moment points from the negative south pole to the positive north pole of this tiny magnet.

57

Magnetic force due to non-uniform magnetic field


Magnets are drawn into regions of higher magnetic field. The simplest example of this is the attraction of opposite poles of two magnets. Every magnet produces a magnetic field that is stronger near its poles. If opposite poles of two separate magnets are facing each other, each of the magnets are drawn into the stronger magnetic field near the pole of the other. If like poles are facing each other though, they are repulsed from the larger magnetic field. The Gilbert model almost predicts the correct mathematical form for this force and is easier to understand qualitatively. For if a magnet is placed in a uniform magnetic field then both poles will feel the same magnetic force but in opposite directions, since they have opposite magnetic charge. But, when a magnet is placed in the non-uniform field, such as that due to another magnet, the pole experiencing the large magnetic field will experience the large force and there will be a net force on the magnet. If the magnet is aligned with the magnetic field, corresponding to two magnets oriented in the same direction near the poles, then it will be drawn into the larger magnetic field. If it is oppositely aligned, such as the case of two magnets with like poles facing each other, then the magnet will be repelled from the region of higher magnetic field. In the physically correct Ampre model, there is also a force on a magnetic dipole due to a non-uniform magnetic field, but this is due to Lorentz forces on the current loop that makes up the magnetic dipole. The force obtained in the case of a current loop model is , where the gradient is the change of the quantity mB per unit distance and the direction is that of maximum increase of mB. To understand this equation, note that the dot product mB = mBcos(), where m and B represent the magnitude of the m and B vectors and is the angle between them. If m is in the same direction as B then the dot product is positive and the gradient points 'uphill' pulling the magnet into regions of higher B-field (more strictly larger mB). This equation is strictly only valid for magnets of zero size, but is often a good approximation for not too large magnets. The magnetic force on larger magnets is determined by dividing them into smaller regions having their own m then summing up the forces on each of these regions.

Gilbert Model
The Gilbert model assumes that the magnetic forces between magnets are due to magnetic charges near the poles. While physically incorrect, this model produces good approximations that work even close to the magnet when the magnetic field becomes more complicated, and more dependent on the detailed shape and magnetization of the magnet than just the magnetic dipole contribution. Formally, the field can be expressed as a multipole expansion: A dipole field, plus a quadrupole field, plus an octopole field, etc. in the Ampre model, but this can be very cumbersome mathematically.

Force between magnets

58

Calculating the magnetic force


Calculating the attractive or repulsive force between two magnets is, in the general case, an extremely complex operation, as it depends on the shape, magnetization, orientation and separation of the magnets. The Gilbert model does depend on some knowledge of how the 'magnetic charge' is distributed over the magnetic poles. It is only truly useful for simple configurations even then. Fortunately, this restriction covers many useful cases. Force between two magnetic poles If both poles are small enough to be represented as a single points then they can be considered to be point magnetic charges. Classically, the force between two magnetic poles is given by:[3]

where F is force (SI unit: newton) qm1 and qm2 are the magnitudes of magnetic poles (SI unit: ampere-meter) is the permeability of the intervening medium (SI unit: tesla meter per ampere, henry per meter or newton per ampere squared) r is the separation (SI unit: meter). The pole description is useful to practicing magneticians who design real-world magnets, but real magnets have a pole distribution more complex than a single north and south. Therefore, implementation of the pole idea is not simple. In some cases, one of the more complex formulas given below will be more useful. Force between two nearby magnetized surfaces of area A The mechanical force between two nearby magnetized surfaces can be calculated with the following equation. The equation is valid only for cases in which the effect of fringing is negligible and the volume of the air gap is much smaller than that of the magnetized material:[4][5]

where: A is the area of each surface, in m2 H is their magnetizing field, in A/m. 0 is the permeability of space, which equals 4107Tm/A B is the flux density, in T

Force between magnets Force between two bar magnets The force between two identical cylindrical bar magnets placed end to end is approximately:[4]

59

where B0 is the magnetic flux density very close to each pole, in T, A is the area of each pole, in m2, L is the length of each magnet, in m, R is the radius of each magnet, in m, and x is the separation between the two magnets, in m relates the flux density at the pole to the magnetization of the magnet. Note that all these formulations are based on the Gilbert's model, which is usable in relatively great distances. In other models, (e.g., Ampre's model) use a more complicated formulation that sometimes cannot be solved analytically. In these cases, numerical methods must be used. Force between two cylindrical magnets For two cylindrical magnets with radius , and height , with their magnetic dipole aligned, the force can be well approximated (even at distances of the order of ) by,[6]
Field of two repelling cylindrical bar magnets Field of two attracting cylindrical bar magnets

Where to

is the magnetization of the magnets and

is the distance between them. In disagreement to the is related

statement in the previous section, a measurement of the magnetic flux density very close to the magnet by the formula

The effective magnetic dipole can be written as

Where

is the volume of the magnet. For a cylinder this is

Force between magnets When the point dipole approximation is obtained,

60

Which matches the expression of the force between two magnetic dipoles.

Ampre model
French scientist Andr Marie Ampre found that the magnetism produced by permanent magnets and the magnetism produced by electromagnets are the same kind of magnetism. Because of that, the strength of a permanent magnet can be expressed in the same terms as that of an electromagnet. The strength of magnetism of an electromagnet that is a flat loop of wire through which a current flows, measured at a distance that is great compared to the size of the loop, is proportional to that current and is proportional to the surface area of that loop. For purpose of expressing the strength of a permanent magnet in same terms as that of an electromagnet, a permanent magnet is thought of as if it contains small current-loops throughout its volume, and then the magnetic strength of that magnet is found to be proportional to the current of each loop (in Ampere), and proportional to the surface of each loop (in square meter), and proportional to the density of current-loops in the material (in units per cubic meter), so the dimension of strength of magnetism of a permanent magnet is Ampere times square meter per cubic meter, is Ampere per meter. That is why Ampere per meter is the correct unit of magnetism, even though these small current loops are not really present in a permanent magnet. The validity of Ampere's model means that it is allowable to think of the magnetic material as if it consists of current-loops, and the total effect is the sum of the effect of each current-loop, and so the magnetic effect of a real magnet can be computed as the sum of magnetic effects of tiny pieces of magnetic material that are at a distance that is great compared to the size of each piece. This is very useful for computing magnetic force-field of a real magnet ; It involves summing a large amount of small forces and you should not do that by hand, but let your computer do that for you ; All that the computer program needs to know is the force between small magnets that are at great distance from each other. In such computations it is often assumed that each (same-size) small piece of magnetic material has an equally strong magnetism, but this is not always true : a magnet that is placed near an other magnet can change the magnetization of that other magnet. For permanent magnets this is usually only a small change, but if you have an electromagnet that consists of a wire wound round an iron core, and you bring a permanent magnet near to that core, then the magnetization of that core can change drastically (for example, if there is no current in the wire, the electromagnet would not be magnetic, but when the permanent magnet is brought near, the core of the electromagnet becomes magnetic). Thus the Ampere model is suitable for computing the magnetic force-field of a permanent magnet, but for electromagnets it can be better to use a magnetic-circuit approach.

Force between magnets

61

Magnetic dipole-dipole interaction


If two or more magnets are small enough or sufficiently distant that their shape and size is not important then both magnets can be modeled as being magnetic dipoles having a magnetic moments m1 and m2. The magnetic field of a magnetic dipole in vector notation is:

field from a perfect dipole, both electric or magnetic.

where B is the field r is the vector from the position of the dipole to the position where the field is being measured r is the absolute value of r: the distance from the dipole is the unit vector parallel to r; m is the (vector) dipole moment 0 is the permeability of free space 3 is the three-dimensional delta function.[7] This is exactly the field of a point dipole, exactly the dipole term in the multipole expansion of an arbitrary field, and approximately the field of any dipole-like configuration at large distances. If the coordinate system is shifted to center it on m1 and rotated such that the z-axis points in the direction of m1 then the previous equation simplifies to[8]

, where the variables r and are measured in a frame of reference with Frames of reference for calculating the forces between two dipoles origin in m1 and oriented such that m1 is at the origin pointing in the z-direction. This frame is called Local coordinates and is shown in the Figure on the right.

Force between magnets The force of one magnetic dipole on another is determined by using the magnetic field of the first dipole given above and determining the force due to the magnetic field on the second dipole using the force equation given above. Using vector notation, the force of a magnetic dipole m1 on the magnetic dipole m2 is:

62

where r is the distance-vector from dipole moment m1 to dipole moment m2, with r=||r||. The force acting on m1 is in opposite direction. As an example the magnetic field for two magnets pointing in the z-direction and aligned on the z-axis and separated by the distance z is: , z-direction. The final formulas are shown next. They are expressed in the global coordinate system,

Notes
[1] Griffiths, David J. (1998). Introduction to Electrodynamics (3rd ed.). Prentice Hall. ISBN0-13-805326-X., section 6.1. [2] The magnetic dipole portion of the magnetic field can be understood as being due to one pair of north/south poles. Higher order terms such as the quadrupole can be considered as due to 2 or more north/south poles ordered such that they have no lower order contribution. For example the quadrupole configuration has no net dipole moment. [3] "Basic Relationships" (http:/ / geophysics. ou. edu/ solid_earth/ notes/ mag_basic/ mag_basic. html). Geophysics.ou.edu. . Retrieved 2009-10-19. [4] "Magnetic Fields and Forces" (http:/ / instruct. tri-c. edu/ fgram/ web/ Mdipole. htm). . Retrieved 2009-12-24. [5] "The force produced by a magnetic field" (http:/ / info. ee. surrey. ac. uk/ Workshop/ advice/ coils/ force. html). . Retrieved 2010-03-09. [6] David Vokoun, Marco Beleggia, Ludek Heller, Petr Sittner, Magnetostatic interactions and forces between cylindrical permanent magnets, Journal of Magnetism and Magnetic Materials, Volume 321, Issue 22, November 2009, Pages 3758-3763, DOI:10.1016/j.jmmm.2009.07.030. [Article http:/ / www. sciencedirect. com/ science/ article/ B6TJJ-4WSRF7C-2/ 2/ 5ede3141fb91e35e83abf6edab5abb94]. Retrieved 02.2009 [7] 3(r) = 0 except at r = (0,0,0), so this term is ignored in multipole expansion. [8] Schill, R. A. (2003). "General relation for the vector magnetic field of a circular current loop: A closer look". IEEE Transactions on Magnetics 39 (2): 961967. Bibcode2003ITM....39..961S. doi:10.1109/TMAG.2003.808597.

References

Lorentz force

63

Lorentz force
In physics, particularly electromagnetism, the Lorentz force is the force on a point charge due to electromagnetic fields. The first derivation of the Lorentz force is commonly attributed to Oliver Heaviside in 1889,[1] although other historians suggest an earlier origin in an 1865 paper by James Clerk Maxwell.[2] Lorentz derived it a few years after Heaviside.

Equation (SI units)


One charged particle
The force F acting on a particle of electric charge q with instantaneous velocity v, due to an external electric field E and magnetic field B, is given by:[3]
Trajectory of a particle with a positive or negative charge q under the influence of a magnetic field B, which is directed perpendicularly out of the screen.

Beam of electrons moving in a circle, due to the presence of a magnetic field. Purple light is emitted along the electron path, due to the electrons colliding with gas molecules in the bulb.

where is the vector cross product. All boldface quantities are vectors. A positively charged particle will be accelerated in the same linear orientation as the E field, but will curve perpendicularly to both the instantaneous velocity vector v and the B field according to the right-hand rule (in detail, if the thumb of the right hand points along v and the index finger along B, then the middle finger points along F). The term qE is called the electric force, while the term qv B is called the magnetic force.[4] According to some definitions, the term "Lorentz force" refers specifically to the formula for the magnetic force,[5] with the total electromagnetic force (including the electric force) given some other (nonstandard) name. This article will not follow this nomenclature: In what follows, the term "Lorentz force" will refer only to the expression for the total force.

Lorentz force The magnetic force component of the Lorentz force manifests itself as the force that acts on a current-carrying wire in a magnetic field. In that context, it is also called the Laplace force.

64

Continuous charge distribution


For a continuous charge distribution in motion, the Lorentz force equation becomes:

where dF is the force on a small piece of the charge distribution with charge dq. If both sides of this equation are divided by the volume of this small piece of the charge distribution dV, the result is:

where f is the force density (force per unit volume) and is the charge density (charge per unit volume). Next, the current density corresponding to the motion of the charge continuum is so the continuous analogue to the equation is[6]

The total force is the volume integral over the charge distribution:

By eliminating and J, using Maxwell's equations, and manipulating using the theorems of vector calculus, this form of the equation can be used to derive the Maxwell stress tensor T, used in General relativity.[6] In terms of the tensor field T and the Poynting vector S, another way to write the Lorentz force (per unit volume) is[6]

where c is the speed of light and denotes the divergence of a tensor field. Rather than the amount of charge and its velocity in electric and magnetic fields, this equation relates the energy flux (flow of energy per unit time per unit distance) in the fields to the force exerted on a charge distribution.

History
Early attempts to quantitatively describe the electromagnetic force were made in the mid-18th century. It was proposed that the force on magnetic poles, by Johann Tobias Mayer and others in 1760, and electrically charged objects, by Henry Cavendish in 1762, obeyed an inverse-square law. However, in both cases the experimental proof was neither complete nor conclusive. It was not until 1784 when Charles-Augustin de Coulomb, using a torsion balance, was able to definitively show through experiment that this was true.[7] Soon after the discovery in 1820 by H. C. rsted that a magnetic needle is acted on by a voltaic current, Andr-Marie Ampre that same year was able to devise through experimentation the formula for the angular dependence of the force between two current elements.[8][9] In all these descriptions, the force was always given in terms of the properties of the objects involved and the distances between them rather than in terms of electric and magnetic fields.[10] The modern concept of electric and magnetic fields first arose in the theories of Michael Faraday, particularly his idea of lines of force, later to be given full mathematical description by Lord Kelvin and James Clerk Maxwell.[11] From a modern perspective it is possible to identify in Maxwell's 1865 formulation of his field equations a form of the Lorentz force equation in relation to electric currents,[2] however, in the time of Maxwell it was not evident how his equations related to the forces on moving charged objects. J. J. Thomson was the first to attempt to derive from Maxwell's field equations the electromagnetic forces on a moving charged object in terms of the object's properties

Lorentz force and external fields. Interested in determining the electromagnetic behavior of the charged particles in cathode rays, Thomson published a paper in 1881 wherein he gave the force on the particles due to an external magnetic field as[1] . Thomson was able to arrive at the correct basic form of the formula, but, because of some miscalculations and an incomplete description of the displacement current, included an incorrect scale-factor of a half in front of the formula. It was Oliver Heaviside, who had invented the modern vector notation and applied them to Maxwell's field equations, that in 1885 and 1889 fixed the mistakes of Thomson's derivation and arrived at the correct form of the magnetic force on a moving charged object.[1][12][13] Finally, in 1892, Hendrik Lorentz derived the modern day form of the formula for the electromagnetic force which includes the contributions to the total force from both the electric and the magnetic fields. Lorentz began by abandoning the Maxwellian descriptions of the ether and conduction. Instead, Lorentz made a distinction between matter and the luminiferous aether and sought to apply the Maxwell equations at a microscopic scale. Using the Heaviside's version of the Maxwell equations for a stationary ether and applying Lagrangian mechanics (see below), Lorentz arrived at the correct and complete form of the force law that now bears his name.[14][15]

65

Trajectories of particles due to the Lorentz force


In many cases of practical interest, the motion in a magnetic field of an electrically charged particle (such as an electron or ion in a plasma) can be treated as the superposition of a relatively fast circular motion around a point called the guiding center and a relatively slow drift of this point. The drift speeds may differ for various species depending on their charge states, masses, or temperatures, possibly resulting in electric currents or chemical separation.

Significance of the Lorentz force


While the modern Maxwell's equations describe how electrically charged particles and currents or moving charged particles give rise to electric and magnetic fields, the Lorentz force law completes that picture by describing the force acting on a moving point charge q in the presence of electromagnetic fields.[3][16] The Lorentz force law describes the effect of E and B upon a point charge, but such electromagnetic forces are not the entire picture. Charged particles are possibly coupled to other forces, notably gravity and nuclear forces. Thus, Maxwell's equations do not stand separate from other physical laws, but are coupled to them via the charge and current densities. The response of a point charge to the Lorentz law is one aspect; the generation of E and B by currents and charges is another. In real materials the Lorentz force is inadequate to describe the behavior of charged particles, both in principle and as a matter of computation. The charged particles in a material medium both respond to the E and B fields and generate these fields. Complex transport equations must be solved to determine the time and spatial response of charges, for example, the Boltzmann equation or the FokkerPlanck equation or the NavierStokes equations. For example, see magnetohydrodynamics, fluid dynamics, electrohydrodynamics, superconductivity, stellar evolution. An entire physical apparatus for dealing with these matters has developed. See for example, GreenKubo relations and Green's function (many-body theory).

Lorentz force

66

Lorentz force law as the definition of E and B


In many textbook treatments of classical electromagnetism, the Lorentz Force Law is used as the definition of the electric and magnetic fields E and B.[17] To be specific, the Lorentz Force is understood to be the following empirical statement: The electromagnetic force on a test charge at a given point and time is a certain function of its charge and velocity, which can be parameterized by exactly two vectors E and B, in the functional form:

If this empirical statement is valid (and, of course, countless experiments have shown that it is), then two vector fields E and B are thereby defined throughout space and time, and these are called the "electric field" and "magnetic field". Note that the fields are defined everywhere in space and time, regardless of whether or not a charge is present to experience the force. In particular, the fields are defined with respect to what force a test charge would feel, if it were hypothetically placed there. Note also that as a definition of E and B, the Lorentz force is only a definition in principle because a real particle (as opposed to the hypothetical "test charge" of infinitesimally-small mass and charge) would generate its own finite E and B fields, which would alter the electromagnetic force that it experiences. In addition, if the charge experiences acceleration, for example, if forced into a curved trajectory by some external agency, it emits radiation that causes braking of its motion. See, for example, Bremsstrahlung and synchrotron light. These effects occur through both a direct effect (called the radiation reaction force) and indirectly (by affecting the motion of nearby charges and currents). Moreover, the electromagnetic force is not in general the same as the net force, due to gravity, electroweak and other forces, and any extra forces would have to be taken into account in a real measurement.

Force on a current-carrying wire


When a wire carrying an electrical current is placed in a magnetic field, each of the moving charges, which comprise the current, experiences the Lorentz force, and together they can create a macroscopic force on the wire (sometimes called the Laplace force). By combining the Lorentz force law above with the definition of electrical current, the following equation results, in the case of a straight, stationary wire:
Right hand rule for a current-carrying wire in a magnetic field B

where is a vector whose magnitude is the length of wire, and whose direction is along the wire, aligned with the direction of conventional current flow I. If the wire is not straight but curved, the force on it can be computed by applying this formula to each infinitesimal segment of wire d, then adding up all these forces by integration. Formally, the net force on a stationary, rigid wire carrying a steady current I is

This is the net force. In addition, there will usually be torque, plus other effects if the wire is not perfectly rigid. One application of this is Ampre's force law, which describes how two current-carrying wires can attract or repel each other, since each experiences a Lorentz force from the other's magnetic field. For more information, see the article: Ampre's force law.

Lorentz force

67

EMF
The magnetic force (q v B) component of the Lorentz force is responsible for motional electromotive force (or motional EMF), the phenomenon underlying many electrical generators. When a conductor is moved through a magnetic field, the magnetic force tries to push electrons through the wire, and this creates the EMF. The term "motional EMF" is applied to this phenomenon, since the EMF is due to the motion of the wire. In other electrical generators, the magnets move, while the conductors do not. In this case, the EMF is due to the electric force (qE) term in the Lorentz Force equation. The electric field in question is created by the changing magnetic field, resulting in an induced EMF, as described by the Maxwell-Faraday equation (one of the four modern Maxwell's equations).[18] Both of these EMF's, despite their different origins, can be described by the same equation, namely, the EMF is the rate of change of magnetic flux through the wire. (This is Faraday's law of induction, see above.) Einstein's theory of special relativity was partially motivated by the desire to better understand this link between the two effects.[18] In fact, the electric and magnetic fields are different faces of the same electromagnetic field, and in moving from one inertial frame to another, the solenoidal vector field portion of the E-field can change in whole or in part to a B-field or vice versa.[19]

Lorentz force and Faraday's law of induction


Given a loop of wire in a magnetic field, Faraday's law of induction states the induced electromotive force (EMF) in the wire is:

where

is the magnetic flux through the loop, B is the magnetic field, (t) is a surface bounded by the closed contour (t), at all at time t, dA is an infinitesimal vector area element of (t) (magnitude is the area of an infinitesimal patch of surface, direction is orthogonal to that surface patch). The sign of the EMF is determined by Lenz's law. Note that this is valid for not only a stationary wire - but also for a moving wire. From Faraday's law of induction (that is valid for a moving wire, for instance in a motor) and the Maxwell Equations, the Lorentz Force can be deduced. The reverse is also true, the Lorentz force and the Maxwell Equations can be used to derive the Faraday Law. Let (t) be the moving wire, moving together without rotation and with constant velocity v and (t) be the internal surface of the wire. The EMF around the closed path (t) is given by:[20]

where

is the electric field and d is an infinitesimal vector element of the contour (t). NB: Both d and dA have a sign ambiguity; to get the correct sign, the right-hand rule is used, as explained in the article Kelvin-Stokes theorem. The above result can be compared with the version of Faraday's law of induction that appears in the modern Maxwell's equations, called here the Maxwell-Faraday equation:

Lorentz force

68

The Maxwell-Faraday equation also can be written in an integral form using the Kelvin-Stokes theorem:.[21] So we have, the Maxwell Faraday equation:

and the Faraday Law,

The two are equivalent if the wire is not moving. Using the Leibniz integral rule and that div B = 0, results in,

and using the Maxwell Faraday equation,

since this is valid for any wire position it implies that,

Faraday's law of induction holds whether the loop of wire is rigid and stationary, or in motion or in process of deformation, and it holds whether the magnetic field is constant in time or changing. However, there are cases where Faraday's law is either inadequate or difficult to use, and application of the underlying Lorentz force law is necessary. See inapplicability of Faraday's law. If the magnetic field is fixed in time and the conducting loop moves through the field, the flux magnetic flux B linking the loop can change in several ways. For example, if the B-field varies with position, and the loop moves to a location with different B-field, B will change. Alternatively, if the loop changes orientation with respect to the B-field, the B dA differential element will change because of the different angle between B and dA, also changing B. As a third example, if a portion of the circuit is swept through a uniform, time-independent B-field, and another portion of the circuit is held stationary, the flux linking the entire closed circuit can change due to the shift in relative position of the circuit's component parts with time (surface (t) time-dependent). In all three cases, Faraday's law of induction then predicts the EMF generated by the change in B. Note that the Maxwell Faraday's equation implies that the Electric Field E is non conservative when the Magnetic Field B varies in time, and is not expressible as the gradient of a scalar field, and not subject to the gradient theorem since its rotational is not zero. See also.[20][22]

Lorentz force in terms of potentials


The E and B fields can be replaced by the magnetic vector potential A and (scalar) electrostatic potential by

where is the gradient, is the divergence, is the curl. The force becomes

and using an identity for the triple product simplifies to

Lorentz force

69

Lorentz force and Lagrangian mechanics


The Lagrangian for a charged particle of mass m and charge q in an electromagnetic field equivalently describes the dynamics of the particle in terms of its energy, rather than the force exerted on it. The classical expression is given by:[23]

where A and are the potential fields as above. Using Lagrange's equations, the equation for the Lorentz force can be obtained.
Derivation of Lorentz force (SI units) For an A field, a particle miving with velocity v = has potential momentum particle's potential energy is The total potential energy is then: . , so its potential energy is . For a field, the

and the kinetic energy is:

hence the Lagrangian:

Lagrange's equations are

(same for y and z). So calculating the partial derivatives:

equating and simplifying:

and similarly for the y and z directions. Hence the force equation is:

The potential energy depends on the velocity of the particle, so the force is velocity dependent, so it is not conservative.

Lorentz force

70

Equation (cgs units)


The above-mentioned formulae use SI units which are the most common among experimentalists, technicians, and engineers. In cgs-Gaussian units, which are somewhat more common among theoretical physicists, one has instead

where c is the speed of light. Although this equation looks slightly different, it is completely equivalent, since one has the following relations:

where 0 is the vacuum permittivity and 0 the vacuum permeability. In practice, the subscripts "cgs" and "SI" are always omitted, and the unit system has to be assessed from context.

Relativistic form of the Lorentz force


Because the electric and magnetic fields are dependent on the velocity of an observer, the relativistic form of the Lorentz force law can best be exhibited starting from a coordinate-independent expression for the electromagnetic and magnetic fields,[24] , and an arbitrary time-direction, , where

and

is a space-time plane (bivector), which has six degrees of freedom corresponding to boosts (rotations in space-time planes) and rotations (rotations in space-space planes). The dot product with the vector pulls a vector from the translational part, while the wedge-product creates a space-time trivector, whose dot product with the volume element (the dual above) creates the magnetic field vector from the spatial rotation part. Only the parts of the above two formulas perpendicular to gamma are relevant. The relativistic velocity is given by the (time-like) changes in a time-position vector , where

(which shows our choice for the metric) and the velocity is

Then the Lorentz force law is simply (note that the order is important)

Covariant form of the Lorentz force


Field tensor Using the metric signature (-1,1,1,1), The Lorentz force for a charge q can be written in covariant form:

where p is the four-momentum, defined as: the proper time of the particle, F the contravariant electromagnetic tensor

Lorentz force

71

and U is the covariant 4-velocity of the particle, defined as:

where

is the Lorentz factor defined above.

The fields are transformed to a frame moving with constant relative velocity by: where is the Lorentz transformation tensor.

Translation to vector notation


The = 1 component (x-component) of the force is

Substituting the components of the covariant electromagnetic tensor F yields

Using the components of covariant four-velocity yields

The calculation for = 2, 3 (force components in the y and z directions) yields similar results, so collecting the 3 equations into one:

which is the Lorentz force.

Applications
The Lorentz force occurs in many devices, including: Cyclotrons and other circular path particle accelerators Mass spectrometers Velocity Filters Magnetrons

In its manifestation as the Laplace force on an electric current in a conductor, this force occurs in many devices including:
Electric motors Railguns Linear motors Loudspeakers Magnetoplasmadynamic thrusters Electrical generators Homopolar generators Linear alternators

Lorentz force

72

References
The numbered references refer in part to the list immediately below. Feynman, Richard Phillips; Leighton, Robert B.; Sands, Matthew L. (2006). The Feynman lectures on physics (3 vol.). Pearson / Addison-Wesley. ISBN0-8053-9047-2: volume 2. Griffiths, David J. (1999). Introduction to electrodynamics (3rd ed.). Upper Saddle River, [NJ.]: Prentice-Hall. ISBN0-13-805326-X Jackson, John David (1999). Classical electrodynamics (3rd ed.). New York, [NY.]: Wiley. ISBN0-471-30932-X Serway, Raymond A.; Jewett, John W., Jr. (2004). Physics for scientists and engineers, with modern physics. Belmont, [CA.]: Thomson Brooks/Cole. ISBN0-534-40846-X Srednicki, Mark A. (2007). Quantum field theory [25]. Cambridge, [England] ; New York [NY.]: Cambridge University Press. ISBN978-0-521-86449-7

Footnotes
[1] Oliver Heaviside By Paul J. Nahin, p120 (http:/ / books. google. com/ books?id=e9wEntQmA0IC& pg=PA120) [2] Huray, Paul G. (2009). Maxwell's Equations (http:/ / books. google. com/ books?id=0QsDgdd0MhMC& pg=PA22#v=onepage& q& f=false). Wiley-IEEE. p.22. ISBN0470542764. . [3] See Jackson page 2. The book lists the four modern Maxwell's equations, and then states, "Also essential for consideration of charged particle motion is the Lorentz force equation, F = q ( E+ v B ), which gives the force acting on a point charge q in the presence of electromagnetic fields." See Griffiths page 204. For example, see the website of the "Lorentz Institute": \ (http:/ / ilorentz. org/ history/ lorentz/ lorentz. html), or Griffiths. Introduction to Electrodynamics (3rd Edition), D.J. Griffiths, Pearson Education, Dorling Kindersley, 2007, ISBN 81-7758-293-3 Meyer, Herbert W. (1972). A History of Electricity and Magnetism. Norwalk, CT: Burndy Library. pp.3031. ISBN026213070X. Verschuur, Gerrit L. (1993). Hidden Attraction : The History And Mystery Of Magnetism. New York: Oxford University Press. pp.7879. ISBN0195064887. [9] Darrigol, Olivier (2000). Electrodynamics from Ampre to Einstein. Oxford, [England]: Oxford University Press. pp.9, 25. ISBN0-198-50593-0 [10] Verschuur, Gerrit L. (1993). Hidden Attraction : The History And Mystery Of Magnetism. New York: Oxford University Press. p.76. ISBN0195064887. [11] Darrigol, Olivier (2000). Electrodynamics from Ampre to Einstein. Oxford, [England]: Oxford University Press. pp.126131, 139144. ISBN0-198-50593-0 [12] Darrigol, Olivier (2000). Electrodynamics from Ampre to Einstein. Oxford, [England]: Oxford University Press. pp.200, 429430. ISBN0-198-50593-0 [13] Heaviside, Oliver. "On the Electromagnetic Effects due to the Motion of Electrification through a Dielectric" (http:/ / en. wikisource. org/ wiki/ Motion_of_Electrification_through_a_Dielectric). Philosophical Magazine, April 1889, p. 324. . [14] Darrigol, Olivier (2000). Electrodynamics from Ampre to Einstein. Oxford, [England]: Oxford University Press. p.327. ISBN0-198-50593-0 [15] Whittaker, E. T. (1910). A History of the Theories of Aether and Electricity: From the Age of Descartes to the Close of the Nineteenth Century (http:/ / books. google. com/ books?id=CGJDAAAAIAAJ& printsec=frontcover#v=onepage& q& f=false). Longmans, Green and Co.. pp.420423. ISBN1143012089. . [16] See Griffiths page 326, which states that Maxwell's equations, "together with the [Lorentz] force law...summarize the entire theoretical content of classical electrodynamics". [17] See, for example, Jackson p777-8. [18] See Griffiths pages 3013. [19] Tai L. Chow (2006). Electromagnetic theory (http:/ / books. google. com/ ?id=dpnpMhw1zo8C& pg=PA153& dq=isbn=0763738271). Sudbury MA: Jones and Bartlett. p.395. ISBN0-7637-3827-1. . [20] Landau, L. D., Lifshits, E. M., & Pitaevski, L. P. (1984). Electrodynamics of continuous media; Volume 8 [[Course of Theoretical Physics (http:/ / worldcat. org/ search?q=0750626348& qt=owc_search)]] (Second ed.). Oxford: Butterworth-Heinemann. p.63 (49 pp. 205207 in 1960 edition). ISBN0750626348. . [4] [5] [6] [7] [8] [21] Roger F Harrington (2003). Introduction to electromagnetic engineering (http:/ / books. google. com/ ?id=ZlC2EV8zvX8C& pg=PA57& dq="faraday's+ law+ of+ induction"). Mineola, NY: Dover Publications. p.56. ISBN0486432416. . [22] M N O Sadiku (2007). Elements of elctromagnetics (http:/ / books. google. com/ ?id=w2ITHQAACAAJ& dq=isbn:0-19-530048-3) (Fourth ed.). NY/Oxford: Oxford University Press. p.391. ISBN0-19-530048-3. .

Lorentz force
[23] Classical Mechanics (2nd Edition), T.W.B. Kibble, European Physics Series, Mc Graw Hill (UK), 1973, ISBN 07-084018-0. [24] Hestenes, David. "SpaceTime Calculus" (http:/ / geocalc. clas. asu. edu/ html/ STC. html). . [25] http:/ / books. google. com/ ?id=5OepxIG42B4C& pg=PA315& dq=isbn=9780521864497

73

External links
Interactive Java tutorial on the Lorentz force (http://www.magnet.fsu.edu/education/tutorials/java/ lorentzforce/index.html) National High Magnetic Field Laboratory Lorentz force (demonstration) (http://www.youtube.com/watch?v=mxMMqNrm598) Faraday's law: Tankersley and Mosca (http://www.nadn.navy.mil/Users/physics/tank/Public/FaradaysLaw. pdf) Notes from Physics and Astronomy HyperPhysics at Georgia State University (http://hyperphysics.phy-astr. gsu.edu/HBASE/hframe.html); see also home page (http://hyperphysics.phy-astr.gsu.edu/HBASE/hframe. html) Interactive Java applet on the magnetic deflection of a particle beam in a homogeneous magnetic field (http:// chair.pa.msu.edu/applets/Lorentz/a.htm) by Wolfgang Bauer

74

Microscopic Origins of Magnetism


Magnetochemistry
Magnetochemistry is concerned with the magnetic properties of chemical compounds. Magnetic properties arise from the spin and orbital angular momentum of the electrons contained in a compound. Compounds are diamagnetic when they contain no unpaired electrons. Molecular compounds that contain one or more unpaired electrons are paramagnetic. The magnitude of the paramagnetism is expressed as an effective magnetic moment, eff. For first-row transition metals the magnitude of eff is, to a first approximation, a simple function of the number of unpaired electrons, the spin-only formula. In general, spin-orbit coupling causes eff to deviate from the spin-only formula. For the heavier transition metals, lanthanides and actinides, spin-orbit coupling cannot be ignored. Exchange interaction can occur in clusters and infinite lattices, resulting in ferromagnetism, antiferromagnetism or ferrimagnetism depending on the relative orientations of the individual spins.

Magnetic susceptibility
The primary measurement in magnetochemistry is magnetic susceptibility. This measures the strength of interaction on placing the substance in a magnetic field. The volume magnetic susceptibility, represented by the symbol is defined by the relationship

where,

is the magnetization of the material (the magnetic dipole moment per unit volume), measured in amperes is the magnetic field strength, also measured in amperes per meter. Susceptibility is a

per meter ( SI units), and

dimensionless quantity. For chemical applications the molar magnetic susceptibility (mol) is the preferred quantity. It is measured in m3mol1 (SI) or cm3mol1 (CGS) and is defined as

where is the density in kgm3 (SI) or gcm3 (CGS) and M is molar mass in kgmol1 (SI) or gmol1 (CGS). A variety of methods are available for the measurement of magnetic susceptibility. With the Gouy balance the weight change of the sample is measured with an analytical balance when the sample is placed in a homogeneous magnetic field. The measurements are calibrated against a known standard, such as mercury cobalt thiocyanate, HgCo(NCS)4. Calibration removes the need to know the density of the sample. Variable temperature measurements can be made by placing the sample in a cryostat between the pole pieces of the magnet.[1]

Schematic diagram of Gouy balance

The Evans balance.[2] is a torsion balance which uses a sample in a fixed position and a variable secondary magnet to bring the magnets back to their initial position. It, too, is calibrated against HgCo(NCS)4. With a Faraday balance the sample is placed in a magnetic field of constant gradient, and weighed on a torsion balance. This method can yield information on magnetic anisotropy.[3] SQUID is a very sensitive magnetometer. For substances in solution NMR may be used to measure susceptibility.[4][5]

Magnetochemistry

75

Types of magnetic behaviour


When an isolated atom is placed in a magnetic field there is an interaction because each electron in the atom behaves like a magnet, that is, the electron has a magnetic moment. There are two type of interaction. 1. Diamagnetism. Every electron is paired with another electron in the same atomic orbital. The moments of the two electrons cancel each other out, so the atom has no net magnetic moment. When placed in a magnetic field the atom becomes magnetically polarized, that is, it develops an induced magnetic moment. The force of the interaction tends to push the atom out of the magnetic field. By convention diamagnetic susceptibility is given a negative sign. 2. Paramagnetism. At least one electron is not paired with another. The atom has a permanent magnetic moment. When placed into a magnetic field, the atom is attracted into the field. By convention paramagnetic susceptibility is given a positive sign. When the atom is present in a chemical compound its magnetic behaviour is modified by its chemical environment. Measurement of the magnetic moment can give useful chemical information. In certain crystalline materials individual magnetic moments may be aligned with each other (magnetic moment has both magnitude and direction). This gives rise to ferromagnetism, antiferromagnetism or ferrimagnetism. These are properties of the crystal as a whole, of little bearing on chemical properties.

Diamagnetism
Diamagnetism is a universal property of chemical compounds, because all chemical compounds contain electron pairs. A compound in which there are no unpaired electrons is said to be diamagnetic. The effect is weak because it depends on the magnitude of the induced magnetic moment. It depends on the number of electron pairs and the chemical nature of the atoms to which they belong. This means that the effects are additive, and a table of "diamagnetic contributions", or Pascal's constants, can be put together.[6][7][8] With paramagnetic compounds the observed susceptibility can be adjusted by adding to it the so-called diamagnetic correction, which is the diamagnetic susceptibility calculated with the values from the table.[9]

Paramagnetism
Mechanism and temperature dependence
A metal ion with a single unpaired electron, such as Cu2+, in a coordination complex provides the simplest illustration of the mechanism of paramagnetism. The individual metal ions are kept far apart by the ligands, so that there is no magnetic interaction between them. The system is said to be magnetically dilute. The magnetic dipoles of the atoms point in random directions. When a magnetic field is applied, first-order Zeeman splitting occurs. Atoms with spins aligned to the field slightly outnumber the atoms with non-aligned spins. In the first-order Zeeman effect the energy difference between the two states is proportional to the applied field strength. Denoting the energy difference as E, the Boltzmann distribution gives the ratio of the two populations as , where k is the Boltzmann constant and T is the

Variation of magnetic susceptibility with temperature

temperature in kelvins. In most cases E is much smaller than kT and the exponential can be expanded as 1 E/kT. It follows from the presence of 1/T in this expression that the susceptibility is inversely proportional to temperature.[10]

Magnetochemistry

76

This is known as the Curie law and the proportionality constant, C, is known as the Curie constant, whose value, for molar susceptibility, is calculated as[11]

where N is the Avagadro constant, g is the Land g-factor, and B is the Bohr magneton. In this treatment it has been assumed that the electronic ground state is not degenerate, that the magnetic susceptibility is due only to electron spin and that only the ground state is thermally populated. While some substances obey the Curie law, others obey the Curie-Weiss law.

Tc is the Curie temperature. The Curie-Weiss law will apply only when the temperature is well above the Curie temperature. At temperatures below the Curie temperature the substance may become ferromagnetic. More complicated behaviour is observed with the heavier transition elements.

Effective magnetic moment


When the Curie law is obeyed, the product of molar susceptibility and temperature is a constant. The effective magnetic moment, eff is then defined[12] as Where C has CGS units cm3 mol-1 K, eff is

Where C has SI units m3 mol-1 K, eff is

The quantity eff is effectively dimensionless, but is often stated as in units of Bohr magneton (B)[12]. For substances that obey the Curie law, the effective magnetic moment is independent of temperature. For other substances eff is temperature dependent, but the dependence is small if the Curie-Weiss law holds and the Curie temperature is low.

Temperature independent paramagnetism


Compounds which are expected to be diamagnetic may exhibit this kind of weak paramagnetism. It arises from a second-order Zeeman effect in which additional splitting, proportional to the square of the field strength, occurs. It is difficult to observe as the compound inevitably also interacts with the magnetic field in the diamagnetic sense. Nevertheless, data are available for the permanganate ion.[13] It is easier to observe in compounds of the heavier elements, such as uranyl compounds.

Magnetochemistry

77

Exchange interactions
Exchange interactions occur when the substance is not magnetically dilute and there are interactions between individual magnetic centres. One of the simplest systems to exhibit the result of exchange interactions is crystalline copper(II) acetate, Cu2(OAc)4(H2O)2. As the formula indicates, it contains two copper(II) ions. The Cu2+ ions are held together by four acetate ligands, each of which binds to both copper ions. Each Cu2+ ion has a d9 electronic configuration, and so should have one unpaired electron. If there were a covalent bond between the copper ions, the electrons would pair up and the compound would be diamagnetic. Instead, there is an exchange interaction in which the spins of the unpaired electrons become partially aligned to each other. In fact two states are created, one with spins parallel and the other with spins opposed. The energy difference between the two states is so small their populations vary significantly with temperature. In consequence the magnetic moment varies with temperature in a sigmoidal pattern. The state with spins opposed has lower energy, so the interaction can be classed as antiferromagnetic in this case.[14] It is believed that this is an example of superexchange, mediated by the oxygen and carbon atoms of the acetate ligands.[15] Other dimers and clusters exhibit exchange behaviour.[16] Exchange interactions can act over infinite chains in one dimension, planes in two dimensions or over a whole crystal in three dimensions. These are examples of long-range magnetic ordering. They give rise to ferromagnetism, antiferromagnetism or ferrimagnetism, depending on the nature and relative orientations of the individual spins.[17]

Copper(II) acetate dihydrate

[14]

Ferrimagnetic ordering in 2 dimensions

Antiferromagnetic ordering in 2 dimensions

Compounds at temperatures below the Curie temperature exhibit long-range magnetic order in the form of ferromagnetism. Another critical temperature is the Nel temperature, below which ferrimagnetism occurs. The hexahydrate of nickel chloride, NiCl26H2O, has a Nel temperature of 8.3 K. The susceptibility is a maximum at this temperature. Below the Nel temperature the susceptibility decreases and the substance becomes antiferromagnetic.[18]

Theoretical calculation for complexes of metal ions


The effective magnetic moment for a compound containing a metal ion with one or more unpaired electrons depends on the total orbital and spin angular momentum of the unpaired electrons, and , respectively. "Total" in this context means "vector sum". In the approximation that the electronic states of the metal ions are determined by Russel-Saunders coupling and that spin-orbit coupling is negligible, the magnetic moment is given by[19]

Magnetochemistry

78

Spin-only formula
Orbital angular momentum is generated when an electron in an orbital of a degenerate set of orbitals is moved to another orbital in the set by rotation. In complexes of high symmetry certain rotations are not possible. In that case the orbital angular momentum is said to be "quenched" and is smaller than might be expected (partial quenching), or zero (complete quenching). There is complete quenching in the following cases. Note that an electron in a degenerate pair of dx2y2 or dz2 orbitals cannot rotate into the other orbital because of symmetry.[20]

Quenched orbital angular momentum


dn Octahedral high-spin low-spin d1 d2 d3 d4 d5 d6 d7 d8 t 6e 2 2g g d9 t 6e 3 2g g t2g3 t2g3eg1 t2g3eg2 t2g6 t2g6eg1 e3t23 e4t23 e1 e2 Tetrahedral

legend: t2g, t2 = (dxy, dxz, dyz). eg, e = (dx2y2, dz2). When orbital angular momentum is completely quenched, and the paramagnetism can be attributed to electron spin alone. The total spin angular momentum is simply half the number of unpaired electrons and the spin-only formula results.

where n is the number of unpaired electrons. The spin-only formula is a good first approximation for high-spin complexes of first-row transition metals.[21]
Ion Number of unpaired electrons 1 1 1 2 2 3 3 3.88 2.83 Spin-only observed moment /B moment /B

Ti3+ V4+ Cu2+ V3+ Ni2+ V2+ Cr3+

1.73

1.73 1.681.78 1.702.20 2.752.85 2.83.5 3.803.90 3.703.90

Magnetochemistry

79
3 3 4 4 5 5 5.92 4.90 4.35.0 3.804.0 4.754.90 5.15.7 5.656.10 5.76.0

Co2+ Mn4+ Cr2+ Fe2+ Mn2+ Fe3+

The small deviations from the spin-only formula may result from the neglect of orbital angular momentum or of spin-orbit coupling. For example, tetrahedral d3, d4, d8 and d9 complexes tend to show larger deviations from the spin-only formula than octahedral complexes of the same ion, because "quenching" of the orbital contribution is less effective in the tetrahedral case.[22]

Low-spin complexes
According to crystal field theory, the d orbitals of a transition metal ion in an octahedal complex are split into two groups in a crystal field. If the splitting is large enough to overcome the energy needed to place electrons in the same orbital, with opposite spin, a low-spin complex will result.
Crystal field diagram for octahedral low-spin d5

Crystal field diagram for octahedral high-spin d5

High and low -spin octahedral complexes


d-count Number of unpaired electrons high-spin d4 d5 d6 d7 4 5 4 3 low-spin 2 1 0 1 Cr2+ , Mn3+ Mn2+, Fe3+ Fe2+, Co3+ Co2+ examples

With one unpaired electron eff values range from 1.8 to 2.5 B and with two unpaired electrons the range is 3.18 to 3.3 B. Note that low-spin complexes of Fe2+ and Co3+ are diamagnetic. Another group of complexes that are diamagnetic are square-planar complexes of d8 ions such as Ni2+ and Rh+ and Au3+.

Magnetochemistry

80

Spin cross-over
When the energy difference between the high-spin and low-spin states is comparable to kT (k is the Boltzmann constant and T the temperature) an equilibrium is established between the spin states, involving what have been called "electronic isomers". Tris-dithiocarbamato iron(III), Fe(S2CNR2)3, is a well-documented example. The effective moment varies from a typical d5 low-spin value of 2.25 B at 80 K to more than 4 B above 300 K.[23]

2nd and 3rd row transition metals


Crystal field splitting is larger for complexes of the heavier transition metals than for the transition metals discussed above. A consequence of this is that low-spin complexes are much more common. Spin-orbit coupling constants, , are also larger and cannot be ignored, even in elementary treatments. The magnetic behaviour has been summarized, as below, together with an extensive table of data.[24]
d-count kT/=0.1 kT/=0 Behaviour with large spin-orbit coupling constant, nd eff eff d1 d2 d3 d4 d5 0.63 1.55 3.88 2.64 1.95 0 1.22 3.88 0 1.73 eff varies with T1/2 eff varies with T, approximately Independent of temperature eff varies with T1/2 eff varies with T, approximately

Lanthanides and actinides


Russel-Saunders coupling, LS coupling, applies to the lanthanide ions, crystal field effects can be ignored, but spin-orbit coupling is not negligible. Consequently spin and orbital angular momenta have to be combined

and the calculated magnetic moment is given by

Magnetic properties of trivalent lanthanide compounds[25]


lanthanide Number of unpaired lectrons calculated moment /B Ce 1 Pr 2 Nd 3 Pm 4 Sm 5 Eu 6 Gd 7 Tb 6 Dy 5 Ho 4 Er 3 Tm 2 Yb 1 Lu 0

2.54

3.58

3.62

2.68

0.85

7.94

9.72

10.65

10.6

9.58

7.56

4.54

observed moment 2.32.5 3.43.6 3.53.6 /B

1.41.7 3.33.5 7.98.0 9.59.8 10.410.6 10.410.7 9.49.6 7.17.5 4.34.9

In actinides spin-orbit coupling is strong and the coupling approximates to j j coupling.

Magnetochemistry

81

This means that it is difficult to calculate the effective moment. For example, uranium(IV), f2, in the complex [UCl6]2 has a measured effective moment of 2.2 B, which includes a contribution from temperature-independent paramagnetism.[26]

Main group elements and organic compounds


Very few compounds of main group elements are paramagnetic. Notable examples include: oxygen, O2; nitric oxide, NO; nitrogen dioxide, NO2 and chlorine dioxide, ClO2. In organic chemistry, compounds with an unpaired electron are said to be free radicals. Free radicals, with some exceptions, are short-lived because one free radical will react rapidly with another, so their magnetic properties are difficult to study. However, if the radicals are well separated from each other in a dilute solution in a solid matrix, at low temperature, they can be studied by electron paramagnetic resonance (EPR). Such radicals are generated by irradiation. Extensive EPR studies have revealed much about electron delocalization in free radicals. The simulated spectrum of the CH3 radical shows hyperfine splitting due to the interaction of the electron with the 3 equivalent hydrogen nuclei, each of which has a spin of 1/2.[27][28]

Simulated EPR spectrum of the CH3 radical

Spin labels are long-lived free radicals which can be inserted into organic molecules so that they can be studied by EPR. [29] For example, the nitroxide MTSL, a functionalized derivative of TEtra Methyl Piperidine Oxide, TEMPO, is used in site-directed spin labeling.

MSTL spin-label

Applications
The gadolinium ion, Gd3+, has the f7 electronic configuration, with all spins parallel. Compounds of the Gd3+ ion are the most suitable for use as a contrast agent for MRI scans.[30] The magnetic moments of gadolinium compounds are larger than those of any transition metal ion. Gadolinium is preferred to other lanthanide ions, some of which have larger effective moments, due to its having a non-degenerate electronic ground state.[31] For many years the nature of oxyhemoglobin, Hb-O2, was highly controversial. It was found experimentally to be diamagnetic. Deoxy-hemoglobin is generally accepted to be a complex of iron in the +2 oxidation state, that is a d6 system with a high-spin magnetic moment near to the spin-only value of 4.9 B. It was proposed that the iron is oxidized and the oxygen reduced to superoxide. Pairing up of electrons from Fe3+ and O2 was then proposed to occur via an exchange mechanism. It has now been shown that in fact the iron(II) changes from high-spin to low-spin when an oxygen molecule donates a pair of electrons to the iron. Whereas in deoxy-hemoglobin the iron atom lies above the plane of the heme, in the low-spin complex the effective ionic radius is reduced and the iron atom lies in the heme plane.[32] Fe(II)Hb + O2 [Fe(II)Hb]O2 (low-spin) Fe(II)Hb (high-spin) + O2 [Fe(III)Hb]O2

This information has an important bearing on research to find artificial oxygen carriers. Compounds of gallium(II) were unknown until quite recently. As the atomic number of gallium is an odd number (31), Ga2+ should have an unpaired electron. It was assumed that it would act as a free radical and have a very short

Magnetochemistry lifetime. The non-existence of Ga(II) compounds was part of the so-called inert pair effect. When salts of the anion with empirical formula such as [GaCl3] were synthesized they were found to be diamagnetic. This implied the formation of a Ga-Ga bond and a dimeric formula, [Ga2Cl6]2.[33]

82

References
[1] Earnshaw, p. 89 [2] Magnetic Susceptibility Balances (http:/ / www. sherwood-scientific. com/ msb/ msbindex. html) [3] O'Connor, C.J. (1982). Lippard, S.J.. ed. Magnetic susceptibility measurements. Progress in Inorganic Chemistry. 29. Wiley. p.203. ISBN978-0-470-16680-2. [4] Evans, D.F. (1959). "The determination of the paramagnetic susceptibility of substances in solution by nuclear magnetic resonance". J. Chem. Soc.,: 20032005. doi:10.1039/JR9590002003. [5] Orchard, p. 15. Earnshshaw, p. 97 [6] Figgis&Lewis, p. 403 [7] Carlin, p. 3 [8] Bain, Gordon A.; Berry , John F. (2008). "Diamagnetic Corrections and Pascals Constants". J. Chem. Educ. 85 (4): 532. Bibcode2008JChEd..85..532B. doi:10.1021/ed085p532. [9] Figgis&Lewis, p.417 [10] Figgis&Lewis, p. 419 [11] Orchard, p. 48 [12] Hoppe, J.I. (1972). "Effective magnetic moment". J. Chem. Educ., 49 (7): 505. Bibcode1972JChEd..49..505H. doi:10.1021/ed049p505. [13] Orchard, p. 53 [14] Lawrence Que (March 2000). Physical methods in bioinorganic chemistry: spectroscopy and magnetism (http:/ / books. google. com/ books?id=KYZ6EXIdjYwC& pg=PA347). University Science Books. pp.345348. ISBN9781891389023. . Retrieved 22 February 2011. [15] Figgis&Lewis, p. 435. Orchard, p. 67 [16] Carlin, sections 5.55.7 [17] Carlin, chapters 6 and 7, pp. 112225 [18] Carin, p. 264 [19] Figgis&Lewis, p. 420 [20] Figgis&Lewis, pp. 424, 432 [21] Figgis&Lewis, p. 406 [22] Figgis&Lewis, Section 3, "Orbital contribution" [23] Orchard, p. 125. Carlin, p. 270 [24] Figgis&Lewis, pp. 443451 [25] Greenwood&Earnshaw p. 1243 [26] Orchard, p. 106 [27] Weil, John A.; Bolton,, James R.; Wertz, John E. (1994). Electron paramagnetic resonance : elementary theory and practical applications. Wiley. ISBN0471572349. [28] Atkins, P. W.; Symons, M. C. R. (1967). The structure of inorganic radicals; an application of electron spin resonance to the study of molecular structure. Elsevier. [29] Berliner, L.J. (1976). Spin labeling : theory and applications I. Academic Press. ISBN0120923505.Berliner, L.J. (1979). Spin labeling II : theory and applications. Academic Press. ISBN0120923521. [30] Krause, W. (2002). Contrast Agents I: Magnetic Resonance Imaging: Pt. 1. Springer. ISBN10: 3540422471. [31] Caravan, Peter; Ellison, Jeffrey J.; McMurry, Thomas J. ; Lauffer, Randall B., Jeffrey J.; McMurry, Thomas J.; Lauffer, Randall B. (1999). "Gadolinium(III) Chelates as MRI Contrast Agents: Structure, Dynamics, and Applications". Chem. Rev. 99 (9): 22932352. doi:10.1021/cr980440x. PMID11749483. [32] Greenwood&Earnshaw, pp. 10991011 [33] Greenwood&Earnshaw, p. 240

Magnetochemistry

83

Bibliography
Carlin, R.L. (1986). Magnetochemistry. Springer. ISBN9783540158165. Earnshaw, Alan (1968). Introduction to Magnetochemistry. Academic Press. Figgis, B.N.; Lewis, J. (1960). "The Magnetochemistry of Complex Compounds". In Lewis. J. and Wilkins. R.G.. Modern Coordination Chemistry. New York: Wiley. Greenwood, N. N.; Earnshaw, A. (1997). Chemistry of the Elements (2nd ed.). Butterworth-Heinemann. ISBN0080379419. Orchard, A.F. (2003). Magnetochemistry. Oxford Chemistry Primers. Oxford University Press. ISBN0198792786. Selwood, P.W. (1943). Magnetochemistry (http://www.archive.org/details/magnetochemistry032763mbp). Interscience Publishers Inc.. Vulfson, Sergey (1998). Molecular Magnetochemistry. Taylor & Francis. ISBN9056995359.

External links
Online available information resources on magnetochemistry (http://www.internetchemie.info/chemistry/ magnetochemistry.htm) Tables of Diagmagnetic Corrections and Pascal's Constants (http://www.jce.divched.org/JCEDLib/ChemInfo/ Inorganic/pascal.html)

Unpaired electron
In chemistry, an unpaired electron is an electron that occupies an orbital of an atom singly, rather than as part of an electron pair. As the formation of electron pairs is often energetically favourable, either in the form of a chemical bond or as a lone pair, unpaired electrons are relatively uncommon in chemistry, because an entity that carries an unpaired electron is usually rather reactive. In organic chemistry they typically only occur briefly during a reaction on an entity called a radical; however, they play an important role in explaining reaction pathways.

Periodic table with elements that have unpaired electrons coloured

Radicals are uncommon in s- and p-block chemistry, since the unpaired electron occupies a valence p orbital or an sp, sp2 or sp3 hybrid orbital. These orbitals are strongly directional and therefore overlap to form strong covalent bonds, favouring dimerisation of radicals. Radicals can be stable if dimerisation would result in a weak bond or the unpaired electrons are stabilised by delocalisation. In contrast, radicals in d- and f-block chemistry are very common. The less directional, more diffuse d and f orbitals, in which unpaired electrons reside, overlap less effectively, form weaker bonds and thus dimerisation is generally disfavoured. These d and f orbitals also have comparatively smaller radial extension, disfavouring overlap to form dimers.[1] The more stable entities with unpaired electrons do exist, e.g. the oxygen-molecule has two unpaired electrons and the nitric oxide molecule has one. According to Hund's rule, the spins of unpaired electrons are aligned parallel and this gives these molecules paramagnetic properties. The most stable examples of unpaired electrons are found on the atoms and ions of lanthanides. The incomplete f-shell of these entities does not interact very strongly with the environment they are in and this prevents them from being paired. The ion with the largest number of unpaired electrons is Gd3+ with seven unpaired electrons.

Unpaired electron

84

References
[1] N. C. Norman (1997). Periodicity and the s- and p-Block Elements. Oxford University Press. p.43. ISBN0-19-855961-5.

Atomic orbital
An atomic orbital is a mathematical function that describes the wave-like behavior of either one electron or a pair of electrons in an atom.[1] This function can be used to calculate the probability of finding any electron of The shapes of the first five atomic orbitals: 1s, 2s, 2px, 2py, and 2pz. The colors show the an atom in any specific region around wave function phase. These are graphs of (x,y,z) functions which depend on the the atom's nucleus. The term may also coordinates of one electron. To see the elongated shape of (x,y,z)2 functions that show refer to the physical region where the probability density more directly, see the graphs of d-orbitals below. electron can be calculated to be, as defined by the particular mathematical form of the orbital.[2] Atomic orbitals are mathematical functions that depend on the coordinates of only one electron. They describe the wave-like nature of this electron in any energy state, and are therefore referred to as wave functions, usually denoted by the symbol (Greek letter psi). They can be the hydrogen-like "orbitals" which are exact solutions to the Schrdinger equation for a hydrogen-like "atom" (i.e., an atom with one electron). Alternatively, atomic orbitals refer to functions that depend on the coordinates of one electron (i.e. orbitals) but are used as starting points for approximating wave functions that depend on the simultaneous coordinates of all the electrons in an atom or molecule. The coordinate systems chosen for atomic orbitals are usually spherical coordinates (r,,) in atoms and cartesians (x,y,z) in poly-atomic molecules. The advantage of spherical coordinates (for atoms) is that an orbital wave function is a product of three factors each dependent on a single coordinate: (r,,) = R(r) () (). Within a visual context, atomic orbitals are the basic building blocks of the introductory pedagogical electron cloud model (derived from the wave mechanics model or atomic orbital model, but using particle concepts in order to visualize the mathematical procedures used to approximate wave functions for atoms with many electrons). Thus (with particle concepts in italics), this model provides a framework for describing the placement of electrons in an atom. In this model, the atom consists of a nucleus surrounded by orbiting electrons. These electrons exist in so called atomic orbitals, which are a set of quantum states of the negatively charged electrons trapped in the electrical field generated by the positively-charged nucleus (which may be weakened by the effect of other electrons, but still remains attractive in sum). The electron cloud model can ultimately be described by quantum mechanics, in which the electrons are more accurately described as standing waves surrounding the nucleus. Atomic orbitals are typically categorized by n, l, and m quantum numbers, which correspond to the electron's energy, angular momentum, and an angular momentum vector component, respectively. Each orbital is defined by a different set of quantum numbers and contains a maximum of two electrons. The simple names s orbital, p orbital, d orbital and f orbital refer to orbitals with angular momentum quantum number l = 0, 1, 2 and 3 respectively. These names indicate the orbital shape and are used to describe the electron configurations. They are derived from the characteristics of their spectroscopic lines: sharp, principal, diffuse, and fundamental, the rest being named in alphabetical order (omittingj).[3][4] The wave function for the electron cloud of a multi-electron atom may be seen as being built up (in approximation) in an electron configuration that is a product of simpler hydrogen-like atomic orbitals. He(x1,y1,z1,x2,y2,z2) 1s(x1,y1,z1) 1s(x2,y2,z2) = 1s2. (This is read as, "The exact wave function depending on the simultaneous coordinates of the two electrons in the helium atom is approximated as a product of two one-s atomic orbitals each

Atomic orbital of which depends on the coordinates of only one electron.") In such a configuration, pairs of electrons are arranged in simple repeating patterns of increasing odd numbers (1,3,5,7..), each of which represents a set of electron pairs with a given energy and angular momentum. The repeating periodicity of the blocks of 2, 6, 10, and 14 elements within sections of the periodic table arises naturally from the total number of electrons which occupy a complete set of s, p, d and f atomic orbitals, respectively. The angular factors of atomic orbitals () () generate s, p, d, etc. functions as real combinations of spherical harmonics Ylm(, ) (where l and m are quantum numbers). There are typically three mathematical forms for the radial functions R(r) which can be chosen as a starting point for the calculation of the properties of atoms and molecules with many electrons. 1. the hydrogen-like atomic orbitals are derived from the exact solution of the Schrdinger Equation for one electron and a nucleus. The part of the function that depends on the distance from the nucleus has nodes (radial nodes) and decays as e(-distance) from the nucleus. 2. The Slater-type orbital (STO) is a form without radial nodes but decays from the nucleus as does the hydrogen-like orbital. 3. The form of the Gaussian type orbital (Gaussians) has no radial nodes and decays as e(-distance squared). Although hydrogen-like orbitals are still used as pedagogical tools, the advent of computers has made STOs preferable for atoms and diatomic molecules since combinations of STOs can replace the nodes in hydrogen-like atomic orbital. Gaussians are typically used in molecules with three or more atoms. Although not as accurate by themselves as STOs, combinations of many Gaussians can attain the accuracy of hydrogen-like orbitals.

85

Introduction
With the development of quantum mechanics, it was found that the orbiting electrons around a nucleus could not be fully described as particles, but needed to be explained by the wave-particle duality. In this sense, the electrons have the following properties:

Wave-like properties
The electrons do not orbit the nucleus in the sense of a planet orbiting the sun, but instead exist as standing waves. The lowest possible energy an electron can take is therefore analogous to the fundamental frequency of a wave on a string. Higher energy states are then similar to harmonics of the fundamental frequency. The electrons are never in a single point location, although the probability of interacting with the electron at a single point can be found from the wave function of the electron.

Particle-like properties
There is always an integer number of electrons orbiting the nucleus. Electrons jump between orbitals in a particle-like fashion. For example, if a single photon strikes the electrons, only a single electron changes states in response to the photon. The electrons retain particle like-properties such as: each wave state has the same electrical charge as the electron particle. Each wave state has a single discrete spin (spin up or spin down).

Visualizing atomic orbitals intuitively


Despite the obvious analogy to planets revolving around the Sun, electrons cannot be described as solid particles. In addition, atomic orbitals do not closely resemble a planet's elliptical path in ordinary atoms. A more accurate analogy might be that of a large and often oddly-shaped "atmosphere" (the electron), distributed around a relatively tiny planet (the atomic nucleus). One difference is that some of an atom's electrons, those in s orbitals, have zero angular momentum, so they cannot in any sense be thought of as moving "around" the nucleus, as a planet does. (A planet

Atomic orbital would need to fall vertically into the Sun and oscillate up and down through it, to be in an orbit with no angular momentum). Other electrons do have varying amounts of angular momentum. Atomic orbitals exactly describe the shape of this "atmosphere" only when a single electron is present in an atom. When more electrons are added to a single atom, the additional electrons tend to more evenly fill in a volume of space around the nucleus so that the resulting collection (sometimes termed the atoms electron cloud [5]) tends toward a generally spherical zone of probability describing where the atoms electrons will be found.

86

History
The term "orbital" was coined by Robert Mulliken in 1932.[6] However, the idea that electrons might revolve around a compact nucleus with definite angular momentum was convincingly argued at least 19 years earlier by Niels Bohr,[7] and the Japanese physicist Hantaro Nagaoka published an orbit-based hypothesis for electronic behavior as early as 1904.[8] Explaining the behavior of these electron "orbits" was one of the driving forces behind the development of quantum mechanics.[9]

Early models
With J.J. Thomson's discovery of the electron in 1897,[10] it became clear that atoms were not the smallest building blocks of nature, but were rather composite particles. The newly discovered structure within atoms tempted many to imagine how the atom's constituent parts might interact with each other. Thomson theorized that multiple electrons revolved in orbit-like rings within a positively-charged jelly-like substance,[11] and between the electron's discovery and 1909, this "plum pudding model" was the most widely-accepted explanation of atomic structure. Shortly after Thomson's discovery, Hantaro Nagaoka, a Japanese physicist, predicted a different model for electronic structure.[8] Unlike the plum pudding model, the positive charge in Nagaoka's "Saturnian Model" was concentrated into a central core, pulling the electrons into circular orbits reminiscent of Saturn's rings. Few people took notice of Nagaoka's work at the time,[12] and Nagaoka himself recognized a fundamental defect in the theory even at its conception, namely that a classical charged object cannot sustain orbital motion because it is accelerating and therefore loses energy due to electromagnetic radiation.[13] Nevertheless, the Saturnian model turned out to have more in common with modern theory than any of its contemporaries.

Bohr atom
In 1909 Ernest Rutherford discovered that the positive half of atoms was tightly condensed into a nucleus,[14] and it became clear from his analysis in 1911 that the plum pudding model could not explain atomic structure. Shortly after, in 1913, Rutherford's postdoctoral student Niels Bohr proposed a new model of the atom, wherein electrons orbited the nucleus with classical periods, but were only permitted to have discrete values of angular momentum, quantized in units h/2.[7] This constraint automatically permitted only certain values of electron energies. The Bohr model of the atom fixed the problem of energy loss from radiation from a ground state (by declaring that there was no state below this), and more importantly explained the origin of spectral lines.

Atomic orbital

87

After Bohr's use of Einstein's explanation of the photoelectric effect to relate energy levels in atoms with the wavelength of emitted light, the connection between the structure of electrons in atoms and the emission and absorption spectra of atoms became an increasingly useful tool in the understanding of electrons in atoms. The most prominent feature of emission and absorption spectra (known experimentally since the middle of the 19th century), was that these atomic spectra contained discrete lines. The significance of the Bohr model was that it related the lines in emission and absorption spectra to the energy differences between the orbits that electrons could take The RutherfordBohr model of the hydrogen around an atom. This was, however, not achieved by Bohr through atom. giving the electrons some kind of wave-like properties, since the idea that electrons could behave as matter waves was not suggested until twelve years later. Still, the Bohr model's use of quantized angular momenta and therefore quantized energy levels was a significant step towards the understanding of electrons in atoms, and also a significant step towards the development of quantum mechanics in suggesting that quantized restraints must account for all discontinuous energy levels and spectra in atoms. With de Broglie's suggestion of the existence of electron matter waves in 1924, and for a short time before the full 1926 Schrdinger equation treatment of hydrogen like atom, a Bohr electron "wavelength" could be seen to be a function of its momentum, and thus a Bohr orbiting electron was seen to orbit in a circle at a multiple of its half-wavelength (this historically incorrect Bohr model is still occasionally taught to students). The Bohr model for a short time could be seen as a classical model with an additional constraint provided by the 'wavelength' argument. However, this period was immediately superseded by the full three-dimensional wave mechanics of 1926. In our current understanding of physics, the Bohr model is called a semi-classical model because of its quantization of angular momentum, not primarily because of its relationship with electron wavelength, which appeared in hindsight a dozen years after the Bohr model was proposed. The Bohr model was able to explain the emission and absorption spectra of hydrogen. The energies of electrons in the n=1, 2, 3, etc. states in the Bohr model match those of current physics. However, this did not explain similarities between different atoms, as expressed by the periodic table, such as the fact that helium (2 electrons), neon (10 electrons), and argon (18 electrons) exhibit similar chemical behavior. Modern physics explains this by noting that the n=1 state holds 2 electrons, the n=2 state holds 8 electrons, and the n=3 state holds 8 electrons (in argon). In the end, this was solved by the discovery of modern quantum mechanics and the Pauli Exclusion Principle.

Modern conceptions and connections to the Heisenberg Uncertainty Principle


Immediately after Heisenberg discovered his uncertainty relation,[15] it was noted by Bohr that the existence of any sort of wave packet implies uncertainty in the wave frequency and wavelength, since a spread of frequencies is needed to create the packet itself.[16] In quantum mechanics, where all particle momenta are associated with waves, it is the formation of such a wave packet which localizes the wave, and thus the particle, in space. In states where a quantum mechanical particle is bound, it must be localized as a wave packet, and the existence of the packet and its minimum size implies a spread and minimal value in particle wavelength, and thus also momentum and energy. In quantum mechanics, as a particle is localized to a smaller region in space, the associated compressed wave packet requires a larger and larger range of momenta, and thus larger kinetic energy. Thus, the binding energy to contain or trap a particle in a smaller region of space, increases without bound, as the region of space grows smaller. Particles cannot be restricted to a geometric point in space, since this would require an infinite particle momentum. In chemistry, Schrdinger, Pauling, Mulliken and others noted that the consequence of Heisenberg's relation was that the electron, as a wave packet, could not be considered to have an exact location in its orbital. Max Born suggested

Atomic orbital that the electron's position needed to be described by a probability distribution which was connected with finding the electron at some point in the wave-function which described its associated wave packet. The new quantum mechanics did not give exact results, but only the probabilities for the occurrence of a variety of possible such results. Heisenberg held that the path of a moving particle has no meaning if we cannot observe it, as we cannot with electrons in an atom. In the quantum picture of Heisenberg, Schrdinger and others, the Bohr atom number n for each orbital became known as an n-sphere in a three dimensional atom and was pictured as the mean energy of the probability cloud of the electron's wave packet which surrounded the atom.

88

Orbital names
Orbitals are given names in the form:

where X is the energy level corresponding to the principal quantum number n, type is a lower-case letter denoting the shape or subshell of the orbital and it corresponds to the angular quantum number l, and y is the number of electrons in that orbital. For example, the orbital 1s2 (pronounced "one ess two") has two electrons and is the lowest energy level (n = 1) and has an angular quantum number of l = 0. In X-ray notation, the principal quantum number is given a letter associated with it. For n = 1, 2, 3, 4, 5, ..., the letters associated with those numbers are K, L, M, N, O, ..., respectively.

Formal quantum mechanical definition


In quantum mechanics, the state of an atom, i.e. the eigenstates of the atomic Hamiltonian, is expanded (see configuration interaction expansion and basis set) into linear combinations of anti-symmetrized products (Slater determinants) of one-electron functions. The spatial components of these one-electron functions are called atomic orbitals. (When one considers also their spin component, one speaks of atomic spin orbitals.) In atomic physics, the atomic spectral lines correspond to transitions (quantum leaps) between quantum states of an atom. These states are labeled by a set of quantum numbers summarized in the term symbol and usually associated with particular electron configurations, i.e., by occupation schemes of atomic orbitals (e.g., 1s2 2s2 2p6 for the ground state of neon -- term symbol: 1S0). This notation means that the corresponding Slater determinants have a clear higher weight in the configuration interaction expansion. The atomic orbital concept is therefore a key concept for visualizing the excitation process associated with a given transition. For example, one can say for a given transition that it corresponds to the excitation of an electron from an occupied orbital to a given unoccupied orbital. Nevertheless, one has to keep in mind that electrons are fermions ruled by the Pauli exclusion principle and cannot be distinguished from the other electrons in the atom. Moreover, it sometimes happens that the configuration interaction expansion converges very slowly and that one cannot speak about simple one-determinant wave function at all. This is the case when electron correlation is large. Fundamentally, an atomic orbital is a one-electron wave function, even though most electrons do not exist in one-electron atoms, and so the one-electron view is an approximation. When thinking about orbitals, we are often given an orbital vision which (even if it is not spelled out) is heavily influenced by this HartreeFock approximation, which is one way to reduce the complexities of molecular orbital theory.

Atomic orbital

89

Hydrogen-like atoms
The simplest atomic orbitals are those that are calculated for systems with a single electron, such as the hydrogen atom. An atom of any other element ionized down to a single electron is very similar to hydrogen, and the orbitals take the same form. In the Schrdinger equation for this system of one negative and one positive particle, the atomic orbitals are the eigenstates of the Hamiltonian operator for the energy. They can be obtained analytically, meaning that the resulting orbitals are products of a polynomial series, and exponential and trigonometric functions. (see hydrogen atom). For atoms with two or more electrons, the governing equations can only be solved with the use of methods of iterative approximation. Orbitals of multi-electron atoms are qualitatively similar to those of hydrogen, and in the simplest models, they are taken to have the same form. For more rigorous and precise analysis, the numerical approximations must be used. A given (hydrogen-like) atomic orbital is identified by unique values of three quantum numbers: n, l, and ml. The rules restricting the values of the quantum numbers, and their energies (see below), explain the electron configuration of the atoms and the periodic table. The stationary states (quantum states) of the hydrogen-like atoms are its atomic orbitals. However, in general, an electron's behavior is not fully described by a single orbital. Electron states are best represented by time-depending "mixtures" (linear combinations) of multiple orbitals. See Linear combination of atomic orbitals molecular orbital method. The quantum number n first appeared in the Bohr model where it determines the radius of each circular electron orbit. In modern quantum mechanics however, n determines the mean distance of the electron from the nucleus; all electrons with the same value of n lie at the same average distance. For this reason, orbitals with the same value of n are said to comprise a "shell". Orbitals with the same value of n and also the same value of l are even more closely related, and are said to comprise a "subshell".

Quantum numbers
Because of the quantum mechanical nature of the electrons around a nucleus, they cannot be described by a location and momentum. Instead, they are described by a set of quantum numbers that encompasses both the particle-like nature and the wave-like nature of the electrons. An atomic orbital is uniquely identified by the values of the three quantum numbers, and each set of the three quantum numbers corresponds to exactly one orbital, but the quantum numbers only occur in certain combinations of values. The quantum numbers, together with the rules governing their possible values, are as follows: The principal quantum number, n, describes the energy of the electron and is always a positive integer. In fact, it can be any positive integer, but for reasons discussed below, large numbers are seldom encountered. Each atom has, in general, many orbitals associated with each value of n; these orbitals together are sometimes called electron shells. The azimuthal quantum number, , describes the orbital angular momentum of each electron and is a non-negative ranges across all (integer) values satisfying the relation , and the n = 2 shell has only orbitals are sometimes collectively integer. Within a shell where n is some integer n0, with , and

. For instance, the n = 1 shell has only orbitals with called a subshell.

. The set of orbitals associated with a particular value of

The magnetic quantum number, , describes the magnetic moment of an electron in an arbitrary direction, and is also always an integer. Within a subshell where is some integer , ranges thus: . The above results may be summarized in the following table. Each cell represents a subshell, and lists the values of available in that subshell. Empty cells represent subshells that do not exist.

Atomic orbital

90

l=0 n=1 n=2 0 n=3 0 n=4 0 n=5 0 ... ...

l=1

l=2

l=3

l=4

...

-1, 0, 1 -1, 0, 1 -2, -1, 0, 1, 2 -1, 0, 1 -2, -1, 0, 1, 2 -3, -2, -1, 0, 1, 2, 3 -1, 0, 1 -2, -1, 0, 1, 2 -3, -2, -1, 0, 1, 2, 3 -4, -3, -2 -1, 0, 1, 2, 3, 4 ... ... ... ... ...

Subshells are usually identified by their

- and

-values.

is represented by its numerical value, but

is

represented by a letter as follows: 0 is represented by 's', 1 by 'p', 2 by 'd', 3 by 'f', and 4 by 'g'. For instance, one may speak of the subshell with and as a '2s subshell'. Each electron also has a spin quantum number, s, which describes the spin of each electron (spin up or spin down). The number s can be +12 or -12. The Pauli exclusion principle states that no two electrons can occupy the same quantum state: every electron in an atom must have a unique combination of quantum numbers.

Shapes of orbitals
Simple pictures showing orbital shapes are intended to describe the angular forms of regions in space where the electrons occupying the orbital are likely to be found. The diagrams cannot, however, show the entire region where an electron can be found, since according to quantum mechanics there is a non-zero probability of finding the electron anywhere in space. Instead the diagrams are approximate representations of boundary or contour surfaces where the probability density |(r,,)|2 has a constant value, chosen so that there is a certain probability (for example 90%) of finding the electron within the contour. Although ||2 as the square of an absolute value is everywhere non-negative, the sign of the wave function (r,,) is often indicated in each subregion of the orbital picture.

Sometimes the function will be graphed to show its phases, rather than the |(r,,)|2 which shows probability density but has no phases (which have been lost in the process of taking the absolute value, since (r,,) is a complex number). |(r,,)|2 orbital graphs tend to have less spherical, thinner lobes than (r,,) graphs, but have the same number of lobes in the same places, and otherwise are recognizable. This article, in order to show wave function phases, shows mostly (r,,) graphs. The lobes can be viewed as interference patterns between the two counter rotating "m" and "-m" modes, with the projection of the orbital onto the xy plane having a resonant "m" wavelengths around the circumference. For each m there are two of these <m>+<-m> and <m>-<-m>. For the case where m=0 the orbital is vertical, counter rotating information is unknown, and the orbital is z-axis symmetric. For the case where l=0 there are no counter rotating modes. There are only radial modes and the shape is spherically symmetric. For any given n, the smaller l is, the more radial nodes there are. Loosely speaking n is energy, l is analogous to eccentricity, and m is orientation. Generally speaking, the number n determines the size and energy of the orbital for a given nucleus: as n increases, the size of the orbital increases. However, in comparing different elements, the higher nuclear charge, Z, of heavier

Cross-section of computed hydrogen atom orbital ((r,,)2) for the 6s (n=6, l=0, m=0) orbital. Note that s orbitals, though spherically symmetrical, have radially placed wave-nodes for n > 1. However, only s orbitals invariably have a center anti-node; the other types never do.

Atomic orbital elements causes their orbitals to contract by comparison to lighter ones, so that the overall size of the whole atom remains very roughly constant, even as the number of electrons in heavier elements (higher Z) increases. Also in general terms, The single s-orbitals ( determines an orbital's shape, and its orientation. However, since some orbitals are also. described by equations in complex numbers, the shape sometimes depends on

91

) are shaped like spheres. For n = 1 the sphere is "solid" (it is most dense at the center

and fades exponentially outwardly), but for n = 2 or more, each single s-orbital is composed of spherically symmetric surfaces which are nested shells (i.e., the "wave-structure" is radial, following a sinusoidal radial component as well). See illustration of a cross-section of these nested shells, at right. The s-orbitals for all n numbers are the only orbitals with an anti-node (a region of high wave function density) at the center of the nucleus. All other orbitals (p, d, f, etc.) have angular momentum, and thus avoid the nucleus (having a wave node at the nucleus). The three p-orbitals for n = 2 have the form of two ellipsoids with a point of tangency at the nucleus (the two-lobed shape is sometimes referred to as a "dumbbell"). The three p-orbitals in each shell are oriented at right angles to each other, as determined by their respective linear combination of values of . Four of the five d-orbitals for n = 3 look similar, each with four pear-shaped lobes, each lobe tangent to two others, and the centers of all four lying in one plane, between a pair of axes. Three of these planes are the xy-, xz-, and yz-planes, and the fourth has the centres on the x and y axes. The fifth and final d-orbital consists of three regions of high probability density: a torus with two pear-shaped regions placed symmetrically on its z axis. There are seven f-orbitals, each with shapes more complex than those of the d-orbitals.
The five d orbitals in (x,y,z) form, with a combination diagram For each s, p, d, f and g set of orbitals, the set of showing how they fit together to fill space around an atomic nucleus. orbitals which composes it forms a spherically symmetrical set of shapes. For non-s orbitals, which have lobes, the lobes point in directions so as to fill space as symmetrically as possible for number of lobes which exist for a set of orientations. For example, the three p orbitals have six lobes which are oriented to each of the six primary directions of 3-D space; for the 5 d orbitals, there are a total of 18 lobes, in which again six point in primary directions, and the 12 additional lobes fill the 12 gaps which exist between each pairs of these 6 primary axes.
2

Additionally, as is the case with the s orbitals, individual p, d, f and g orbitals with n values higher than the lowest possible value, exhibit an additional radial node structure which is reminiscent of harmonic waves of the same type, as compared with the lowest (or fundamental) mode of the wave. As with s orbitals, this phenomenon provides p, d, f, and g orbitals at the next higher possible value of n (for example, 3p orbitals vs. the fundamental 2p), an additional node in each lobe. Still higher values of n further increase the number of radial nodes, for each type of orbital. The shapes of atomic orbitals in one-electron atom are related to 3-dimensional spherical harmonics. These shapes are not unique, and any linear combination is valid, like a transformation to cubic harmonics, in fact it is possible to generate sets where all the d's are the same shape, just like the px, py, and pz are the same shape.[17][18]

Atomic orbital

92

Orbitals table
This table shows all orbital configurations for the real hydrogen-like wave functions up to 7s, and therefore covers the simple electronic configuration for all elements in the periodic table up to radium. graphs are shown with - and + wave function phases shown in two different colors (arbitrarily red and blue). The pz orbital is the same as the p0 orbital, but the px and py are formed by taking linear combinations of the p+1 and p-1 orbitals (which is why they are listed under the m=1 label). Also, the p+1 and p-1 are not the same shape as the p0, since they are pure spherical harmonics.
s (l=0) m=0 s n=1 n=2 n=3 n=4 p (l=1) m=0 pz m=1 px py m=0 dz2 d (l=2) m=1 dxz dyz m=2 dxy d 2 2 x -y m=0 fz3 m=1 f (l=3) m=2 m=3

fxz2 fyz2 fxyz fz(x2-y2) fx(x2-3y2) fy(3x2-y2)

n=5 n=6 n=7 ... ... ... ... ... ... ... ... ... ... ... ... ...

... ... ...

... ... ...

... ... ...

... ... ...

... ... ...

... ... ...

... ... ...

Qualitative understanding of shapes


The shapes of atomic orbitals can be understood qualitatively by considering the analogous case of standing waves on a circular drum.[19] To see the analogy, the mean vibrational displacement of each bit of drum membrane from the equilibrium point over many cycles (a measure of average drum membrane velocity and momentum at that point) must be considered relative to that point's distance from the center of the drum head. If this displacement is taken as being analogous to the probability of finding an electron at a given distance from the nucleus, then it will be seen that the many modes of the vibrating disk form patterns that trace the various shapes of atomic orbitals. The basic reason for this correspondence lies in the fact that the distribution of kinetic energy and momentum in a matter-wave is predictive of where the particle associated with the wave will be. That is, the probability of finding an electron at a given place is also a function of the electron's average momentum at that point, since high electron momentum at a given position tends to "localize" the electron in that position, via the properties of electron wave-packets (see the Heisenberg uncertainty principle for details of the mechanism). This relationship means that certain key features can be observed in both drum membrane modes and atomic orbitals. For example, in all of the modes analogous to s orbitals (the top row in the animated illustration below), it can be seen that the very center of the drum membrane vibrates most strongly, corresponding to the antinode in all s orbitals in an atom. This antinode means the electron is most likely to be at the physical position of the nucleus (which it passes straight through without scattering or striking it), since it is moving (on average) most rapidly at that point, giving it maximal momentum. A mental "planetary orbit" picture closest to the behavior of electrons in s orbitals, all of which have no angular momentum, might perhaps be that of the path of an atomic-sized black hole, or some other imaginary particle which is able to fall with increasing velocity from space directly through the Earth, without stopping or being affected by

Atomic orbital any force but gravity, and in this way falls through the core and out the other side in a straight line, and off again into space, while slowing from the backwards gravitational tug. If such a particle were gravitationally bound to the Earth it would not escape, but would pursue a series of passes in which it always slowed at some maximal distance into space, but had its maximal velocity at the Earth's center (this "orbit" would have an orbital eccentricity of 1.0). If such a particle also had a wave nature, it would have the highest probability of being located where its velocity and momentum were highest, which would be at the Earth's core. In addition, rather than be confined to an infinitely narrow "orbit" which is a straight line, it would pass through the Earth from all directions, and not have a preferred one. Thus, a "long exposure" photograph of its motion over a very long period of time, would show a sphere. In order to be stopped, such a particle would need to interact with the Earth in some way other than gravity. In a similar way, all s electrons have a finite probability of being found inside the nucleus, and this allows s electrons to occasionally participate in strictly nuclear-electron interaction processes, such as electron capture and internal conversion. Below, a number of drum membrane vibration modes are shown. The analogous wave functions of the hydrogen atom are indicated. A correspondence can be considered where the wave functions of a vibrating drum head are for a two-coordinate system (r,) and the wave functions for a vibrating sphere are three-coordinate (r,,). s-type modes

93

Mode

(1s orbital)

Mode

(2s orbital)

Mode

(3s orbital)

None of the other sets of modes in a drum membrane have a central antinode, and in all of them the center of the drum does not move. These correspond to a node at the nucleus for all non-s orbitals in an atom. These orbitals all have some angular momentum, and in the planetary model, they correspond to particles in orbit with eccentricity less than 1.0, so that they do not pass straight through the center of the primary body, but keep somewhat away from it. In addition, the drum modes analogous to p and d modes in an atom show spatial irregularity along the different radial directions from the center of the drum, whereas all of the modes analogous to s modes are perfectly symmetrical in radial direction. The non radial-symmetry properties of non-s orbitals are necessary to localize a particle with angular momentum and a wave nature in an orbital where it must tend to stay away from the central attraction force, since any particle localized at the point of central attraction could have no angular momentum. For these modes, waves in the drum head tend to avoid the central point. Such features again emphasize that the shapes of atomic orbitals are a direct consequence of the wave nature of electrons. p-type modes

Mode

(2p orbital)

Mode

(3p orbital)

Mode

(4p orbital)

d-type modes

Atomic orbital

94

Mode

(3d orbital)

Mode

(4d orbital)

Mode

(5d orbital)

Orbital energy
In atoms with a single electron (hydrogen-like atoms), the energy of an orbital (and, consequently, of any electrons in the orbital) is determined exclusively by . The orbital has the lowest possible energy in the atom. Each successively higher value of has a higher level of energy, but the difference decreases as increases. For high , the level of energy becomes so high that the electron can easily escape from the atom. In single electron atoms, all levels with different within a given are (to a good approximation) degenerate, and have the same energy. [This approximation is broken to a slight extent by the effect of the magnetic field of the nucleus, and by quantum electrodynamics effects. The latter induce tiny binding energy differences especially for s electrons that go nearer the nucleus, since these feel a very slightly different nuclear charge, even in one-electron atoms. See Lamb shift.] In atoms with multiple electrons, the energy of an electron depends not only on the intrinsic properties of its orbital, but also on its interactions with the other electrons. These interactions depend on the detail of its spatial probability distribution, and so the energy levels of orbitals depend not only on but also on . Higher values of are associated with higher values of energy; for instance, the 2p state is higher than the 2s state. When = 2, the increase in energy of the orbital becomes so large as to push the energy of orbital above the energy of the s-orbital in the next higher shell; when = 3 the energy is pushed into the shell two steps higher. The filling of the 3d orbitals does not occur until the 4s orbitals have been filled. The increase in energy for subshells of increasing angular momentum in larger atoms is due to electron-electron interaction effects, and it is specifically related to the ability of low angular momentum electrons to penetrate more effectively toward the nucleus, where they are subject to less screening from the charge of intervening electrons. Thus, in atoms of higher atomic number, the of electrons becomes more and more of a determining factor in their energy, and the principal quantum numbers placement. of electrons becomes less and less important in their energy

The energy sequence of the first 24 subshells (e.g., 1s, 2p, 3d, etc.) is given in the following table. Each cell represents a subshell with and given by its row and column indices, respectively. The number in the cell is the subshell's position in the sequence. For a linear listing of the subshells in terms of increasing energies in multielectron atoms, see the section below.

Atomic orbital

95

1 1 2 2 3 4 4 6 5 9 3 5 8 7 10 13

11 14 17 21

6 12 15 18 22 7 16 19 23 8 20 24

Note: empty cells indicate non-existent sublevels, while numbers in italics indicate sublevels that could exist, but which do not hold electrons in any element currently known.

Electron placement and the periodic table


Several rules govern the placement of electrons in orbitals (electron configuration). The first dictates that no two electrons in an atom may have the same set of values of quantum numbers (this is the Pauli exclusion principle). These quantum numbers include the three that define orbitals, as well as s, or spin quantum number. Thus, two electrons may occupy a single orbital, so long as they have different values of . However, only two electrons, because of their spin, can be associated with each orbital.

Additionally, an electron always tends to fall to the lowest possible energy state. It is possible for it to occupy any orbital so long as it does not violate the Pauli exclusion principle, but if lower-energy orbitals are available, this condition is unstable. The electron will eventually lose energy (by releasing a photon) and drop into the lower orbital. Thus, electrons fill orbitals in the order specified by the energy sequence given above. This behavior is responsible for the structure of the periodic table. The table may be divided into several rows (called 'periods'), numbered starting with 1 at the top. The presently known elements occupy seven periods. If a certain period has number , it consists of elements whose outermost electrons fall in the th shell. Niels Bohr was the first to propose (1923) that the periodicity in the properties of the elements might be explained by the periodic filling of the electron energy levels, resulting in the electronic structure of the atom.[20] The periodic table may also be divided into several numbered rectangular 'blocks'. The elements belonging to a given block have this common feature: their highest-energy electrons all belong to the same -state (but the associated with that -state depends upon the period). For instance, the leftmost two columns constitute the 's-block'. The

Electron atomic and molecular orbitals. The chart of orbitals (left) is arranged by increasing energy (see Madelung rule). Note that atomic orbits are functions of three variables (two angles, and the distance from the nucleus, r). These images are faithful to the angular component of the orbital, but not entirely representative of the orbital as a whole.

Atomic orbital outermost electrons of Li and Be respectively belong to the 2s subshell, and those of Na and Mg to the 3s subshell. The following is the order for filling the "subshell" orbitals, which also gives the order of the "blocks" in the periodic table: 1s, 2s, 2p, 3s, 3p, 4s, 3d, 4p, 5s, 4d, 5p, 6s, 4f, 5d, 6p, 7s, 5f, 6d, 7p The "periodic" nature of the filling of orbitals, as well as emergence of the s, p, d and f "blocks", is more obvious if this order of filling is given in matrix form, with increasing principal quantum numbers starting the new rows ("periods") in the matrix. Then, each subshell (composed of the first two quantum numbers) is repeated as many times as required for each pair of electrons it may contain. The result is a compressed periodic table, with each entry representing two successive elements:
1s 2s 3s 4s 3d 5s 4d 6s (4f) 5d 7s (5f) 6d

96

3d 4d 5d 6d

3d 4d 5d 6d

3d 4d 5d 6d

3d 4d 5d 6d

2p 3p 4p 5p 6p 7p

2p 3p 4p 5p 6p 7p

2p 3p 4p 5p 6p 7p

The number of electrons in an electrically neutral atom increases with the atomic number. The electrons in the outermost shell, or valence electrons, tend to be responsible for an element's chemical behavior. Elements that contain the same number of valence electrons can be grouped together and display similar chemical properties.

Relativistic effects
For elements with high atomic number Z, the effects of relativity become more pronounced, and especially so for s electrons, which move at relativistic velocities as they penetrate the screening electrons near the core of high Z atoms. This relativistic increase in momentum for high speed electrons causes a corresponding decrease in wavelength and contraction of 6s orbitals relative to 5d orbitals (by comparison to corresponding s and d electrons in lighter elements in the same column of the periodic table); this results in 6s valence electrons becoming lowered in energy. Examples of significant physical outcomes of this effect include the lowered melting temperature of mercury (which results from 6s electrons not being available for metal bonding) and the golden color of gold and caesium (which results from narrowing of 6s to 5d transition energy to the point that visible light begins to be absorbed).[21] In the Bohr Model, an electron has a velocity given by , where Z is the atomic number, is the fine-structure constant, and c is the speed of light. In non-relativistic quantum mechanics, therefore, any atom with an atomic number greater than 137 would require its 1s electrons to be traveling faster than the speed of light. Even in the Dirac equation, which accounts for relativistic effects, the wavefunction of the electron for atoms with Z > 137 is oscillatory and unbounded. The significance of element 137, also known as untriseptium, was first pointed out by the physicist Richard Feynman. Element 137 is sometimes informally called feynmanium (symbol Fy). However, Feynman's approximation fails to predict the exact critical value of Z due to the non-point-charge nature of the nucleus and very small orbital radius of inner electrons, resulting in a potential seen by inner electrons which is effectively less than Z. The critical Z value which makes the atom unstable with regard to high-field breakdown of the vacuum and production of electron-positron pairs, does not occur until Z is about 173. These conditions are not seen except transiently in collisions of very heavy nuclei such as lead or uranium in accelerators, where such electron-positron production from these effects has been claimed to be observed. See Extension of the periodic table beyond the seventh period. There are no nodes in relativistic orbital densities, although individual components of the wavefunction will have nodes.[22]

Atomic orbital

97

Transitions between orbitals


Under quantum mechanics, each quantum state has a well-defined energy. When applied to atomic orbitals, this means that each state has a specific energy, and that if an electron is to move between states, the energy difference is also very fixed. Consider two states of the Hydrogen atom: State 1) n=1, l=0, ml=0 and s=+12 State 2) n=2, l=0, ml=0 and s=+12 By quantum theory, state 1 has a fixed energy of E1, and state 2 has a fixed energy of E2. Now, what would happen if an electron in state 1 were to move to state 2? For this to happen, the electron would need to gain an energy of exactly E2 - E1. If the electron receives energy that is less than or greater than this value, it cannot jump from state 1 to state 2. Now, suppose we irradiate the atom with a broad-spectrum of light. Photons that reach the atom that have an energy of exactly E2 - E1 will be absorbed by the electron in state 1, and that electron will jump to state 2. However, photons that are greater or lower in energy cannot be absorbed by the electron, because the electron can only jump to one of the orbitals, it cannot jump to a state between orbitals. The result is that only photons of a specific frequency will be absorbed by the atom. This creates a line in the spectrum, known as an absorption line, which corresponds to the energy difference between states 1 and 2. The atomic orbital model thus predicts line spectra, which are observed experimentally. This is one of the main validations of the atomic orbital model. The atomic orbital model is nevertheless an approximation to the full quantum theory, which only recognizes many electron states. The predictions of line spectra are qualitatively useful but are not quantitatively accurate for atoms and ions other than those containing only one electron.

References
[1] Orchin, Milton; Macomber, Roger S.; Pinhas, Allan; Wilson, R. Marshall (2005). Atomic Orbital Theory (http:/ / media. wiley. com/ product_data/ excerpt/ 81/ 04716802/ 0471680281. pdf). . [2] Daintith, J. (2004). Oxford Dictionary of Chemistry. New York: Oxford University Press. ISBN0-19-860918-3. [3] Griffiths, David (1995). Introduction to Quantum Mechanics. Prentice Hall. pp.190191. ISBN0-13-124405-1. [4] Levine, Ira (2000). Quantum Chemistry (5 ed.). Prentice Hall. pp.144145. ISBN0-13-685512-1. [5] Feynman, Richard; Leighton, Robert B.; Sands, Matthew (2006). The Feynman Lectures on Physics -The Definitive Edition, Vol 1 lect 6. Pearson PLC, Addison Wesley. p.11. ISBN0-8053-9046-4. [6] Mulliken, Robert S. (July 1932). "Electronic Structures of Polyatomic Molecules and Valence. II. General Considerations". Physical Review 41 (1): 4971. Bibcode1932PhRv...41...49M. doi:10.1103/PhysRev.41.49. [7] Bohr, Niels (1913). "On the Constitution of Atoms and Molecules" (http:/ / www. chemteam. info/ Chem-History/ Bohr/ Bohr-1913a. html). Philosophical Magazine 26 (1): 476. . [8] Nagaoka, Hantaro (May 1904). "Kinetics of a System of Particles illustrating the Line and the Band Spectrum and the Phenomena of Radioactivity" (http:/ / www. chemteam. info/ Chem-History/ Nagaoka-1904. html). Philosophical Magazine 7: 445455. . [9] Bryson, Bill (2003). A Short History of Nearly Everything. Broadway Books. pp.141143. ISBN0-7679-0818-X. [10] Thomson, J. J. (1897). "Cathode rays". Philosophical Magazine 44: 293. [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" (http:/ / www. chemteam. info/ Chem-History/ Thomson-Structure-Atom. html) (extract of paper). Philosophical Magazine Series 6 7 (39). doi:10.1080/14786440409463107. . [12] Rhodes, Richard (1995). The Making of the Atomic Bomb (http:/ / books. google. com/ ?id=aSgFMMNQ6G4C& printsec=frontcover& dq=making+ of+ the+ atomic+ bomb#v=onepage& q& f=false). Simon & Schuster. pp.5051. ISBN9780684813783. . [13] Nagaoka, Hantaro (May 1904). "Kinetics of a System of Particles illustrating the Line and the Band Spectrum and the Phenomena of Radioactivity" (http:/ / www. chemteam. info/ Chem-History/ Nagaoka-1904. html). Philosophical Magazine 7: 446. . [14] Geiger, H.; Marsden, E. (1909). "On a Diffuse Reflection of the -Particles" (http:/ / www. chemteam. info/ Chem-History/ GM-1909. html). Proceedings of the Royal Society, Series A 82 (557): 495500. Bibcode1909RSPSA..82..495G. doi:10.1098/rspa.1909.0054. . [15] Heisenberg, W. (March 1927). "ber den anschaulichen Inhalt der quantentheoretischen Kinematik und Mechanik" (http:/ / www. springerlink. com/ content/ t8173612621026q5/ ). Zeitschrift fr Physik A 43 (34): 172198. Bibcode1927ZPhy...43..172H. doi:10.1007/BF01397280. .

Atomic orbital
[16] Bohr, Niels (April 1928). "The Quantum Postulate and the Recent Development of Atomic Theory" (http:/ / www. nature. com/ doifinder/ 10. 1038/ 121580a0). Nature 121 (3050): 580590. Bibcode1928Natur.121..580B. doi:10.1038/121580a0. . [17] Powell, Richard E. (1968). "The five equivalent d orbitals". Journal of Chemical Education 45 (1): 45. Bibcode1968JChEd..45...45P. doi:10.1021/ed045p45. [18] Kimball, George E. (1940). "Directed Valence". The Journal of Chemical Physics 8 (2): 188. Bibcode1940JChPh...8..188K. doi:10.1063/1.1750628. [19] Cazenave, Lions, T., P.; Lions, P. L. (1982). "Orbital stability of standing waves for some nonlinear Schrdinger equations". Communications in Mathematical Physics 85 (4): 549561. Bibcode1982CMaPh..85..549C. doi:10.1007/BF01403504. [20] Bohr, Niels (1923). "ber die Anwendung der Quantumtheorie auf den Atombau. I". Zeitschrift fr Physik 13: 117. Bibcode1923ZPhy...13..117B. doi:10.1007/BF01328209. [21] Lower, Stephen. "Primer on Quantum Theory of the Atom" (http:/ / www. chem1. com/ acad/ webtut/ atomic/ qprimer/ #Q26). . [22] Szabo, Attila (1969). "Contour diagrams for relativistic orbitals". Journal of Chemical Education 46 (10): 678. Bibcode1969JChEd..46..678S. doi:10.1021/ed046p678.

98

Further reading
Tipler, Paul; Llewellyn, Ralph (2003). Modern Physics (4 ed.). New York: W. H. Freeman and Company. ISBN0-7167-4345-0. Scerri, Eric (2007). The Periodic Table, Its Story and Its Significance. New York: Oxford University Press. ISBN978-0-19-530573-9. Levine, Ira (2000). Quantum Chemistry. Upper Saddle River, New Jersey: Prentice Hall. ISBN0-13-685512-1. Griffiths, David (2000). Introduction to Quantum Mechanics (2 ed.). Benjamin Cummings. ISBN978-0131118928. Cohen, Irwin; Bustard, Thomas (1966). "Atomic Orbitals: Limitations and Variations" (http://pubs.acs.org/doi/ pdfplus/10.1021/ed043p187). J. Chem. Educ. 43 (4): 187. Bibcode1966JChEd..43..187C. doi:10.1021/ed043p187.

External links
Guide to atomic orbitals (http://www.chemguide.co.uk/atoms/properties/atomorbs.html) Covalent Bonds and Molecular Structure (http://wps.prenhall.com/wps/media/objects/602/616516/ Chapter_07.html) Animation of the time evolution of an hydrogenic orbital (http://strangepaths.com/atomic-orbital/2008/04/20/ en/) The Orbitron (http://www.shef.ac.uk/chemistry/orbitron/), a visualization of all common and uncommon atomic orbitals, from 1s to 7g Grand table (http://www.orbitals.com/orb/orbtable.htm) Still images of many orbitals David Manthey's Orbital Viewer (http://www.orbitals.com/orb/index.html) renders orbitals with n30 Java orbital viewer applet (http://www.falstad.com/qmatom/) What does an atom look like? Orbitals in 3D (http://www.hydrogenlab.de/elektronium/HTML/ einleitung_hauptseite_uk.html) Atom Orbitals v.1.5 visualization software (http://taras-zavedy.narod.ru/PROGRAMMS/ ATOM_ORBITALS_v_1_5_ENG/Atom_Orbitals_v_1_5_ENG.html)

d electron count

99

d electron count
The d electron count is a chemistry formalism used to describe the electron configuration of the valence electrons of a transition metal center in a coordination complex.[1][2] The d electron count is an effective way to understand the geometry and reactivity of transition metal complexes. The formalism has been incorporated into the two major models used to describe coordination complexes; ligand field theory which is an application of molecular orbital theory to transition metals and crystal field theory which has roots in VSEPR theory.[3]

Standard electron configuration perspective


The electron configuration for transition metals predicted by the simple Aufbau principle and Madelung's rule has serious conflicts with experimental observations for transition metal centers under most ambient conditions. Under most conditions all of the valence electrons of a transition metal center are located in d orbitals while the standard model of electron configuration would predict some of them to be in the pertinent s orbital. The valence of a transition metal center can be described by standard quantum numbers. The Aufbau principle and Madelung's rule would predict for period n that the ns orbitals fill prior to the (n-1)d orbitals. For example the 4s fills before the 3d in period 4. In general chemistry textbooks, a few exceptions are acknowledged with only one electron in the ns orbital in favor of completing a half or whole d shell. The usual explanation is that "half-filled or completely filled subshells are particularly stable arrangements of electrons". An example is chromium whose electron configuration is [Ar] 4s1 3d5 with a half-filled d subshell, although Madelung's rule would predict 4s2 3d4. Similarly copper is [Ar] 4s1 3d10 with a full d subshell, and not [Ar] 4s2 3d9.[4] Matters are further complicated when metal centers are oxidized. Since the (n-1)d shell is predicted to have higher energy than the ns shell, it might be expected that electrons would be removed from the (n-1)d shell first. Experimentally it has been observed that not only are the ns electrons removed first, even for unionized complexes all of the valence electrons are located in the (n-1)d orbitals. There are various hand waving arguments for this phenomenon including that "the ns electrons are farther away from the nuclei and thus ionized first" while ignoring results based on neutral complexes. This poor explanation avoids the basic problems with the standard electron configuration model. The standard electron configuration model assumes a hydrogen-like atom removed from all other atoms. This assumption is only truly relevant for esoteric situations. It is far more common for metal centers to have bonds to other atoms through metallic bonds or covalent bonds. These bonds drastically change the energies of the orbitals for which electron configurations are predicted. Thus for coordination complexes the standard electron configuration formalism is meaningless and the d electron count formalism is a suitable substitute.

d electron count

100

Ligand field perspective


Crystal field theory describes a number of physical phenomena well but does not describe bonding nor offer an explanation for why ns electrons are ionized before (n-1)d electrons. The more recent ligand field theory offers an easy to understand explanation the models phenomenon relatively well. According to the model present by ligand field theory, the ns orbital is involved in bonding to the ligands and forms a strongly bonding orbital which has predominantly ligand character and the correspondingly strong anti-bonding orbital which is unfilled and usually well above the lowest Ligand-Field scheme summarizing -bonding in the octahedral complex [Ti(H2O)6]3+. unoccupied molecular orbital (LUMO). Since the orbitals resulting form the ns orbital are either buried in bonding or elevated well above the valence the ns orbitals are not relevant to the describing the valence. Depending on the geometry of the final complex either all three of the np orbitals or a portions of them are involved in bonding similar to the ns orbitals. The np orbitals if any that remain non-bonding still exceed the valence of the complex. That leaves the (n-1)d orbitals to be involved in some portion of the bonding and in the process also describe the metal complexes valence electrons. The final description of the valence is highly dependent on the complexes geometry. The complexes geometry is in turn highly dependent on the d electron count and character of the associated ligands. For example in the MO diagram provided for the [Ti(H2O)6]3+ the ns orbital (which is placed above (n-1)d in the representation of atomic orbitals (AO)) is used in a linear combination with the ligand orbitals forming a very stable bonding orbital with significant ligand character as well as an unoccupied high energy anti-bonding orbital which is not shown. In this situation the complex geometry is octahedral which means the two of the d orbitals have the proper geometry to be involved in bonding. The other three d orbitals in the basic model do not have significant interactions with the ligands and remain three degenerate non-bonding orbitals. The two orbitals that are involved in bonding form a linear combination with two ligand orbitals with the proper symmetry. This results in two filled bonding orbitals and two orbitals which are usually the lowest unoccupied molecular orbitals (LUMO) or the highest partially filled molecular orbitals a variation on the high occupied molecular orbitals (HOMO).

Tanabe-Sugano diagram
Each of the ten possible d electron counts has an associated Tanabe-Sugano diagram describing gradations of possible ligand field environments a metal center could experience in an octahedral geometry. The Tanabe-Sugano diagram with a small amount of information accurately predicts absorptions in the UV and visible electromagnetic spectrum resulting from d to d orbital electron transitions. It is these d-d transitions, ligand to metal charge transfers (LMCT), or metal to ligand charge transfers (MLCT) that generally give metals complexes their vibrant colors.

d electron count

101

Limitation
It is important to remember that the d electron count is a formalism and describes some complexes better than others. Often it is difficult or impossible to assign electrons and charge to the metal center or a ligand. For a high oxidation state metal center with a 4+ charge or greater it is understood that the true charge separation is much smaller. But referring to the formal oxidation state and d electron count can still be useful when trying to understand the chemistry.

Possible d electron counts


There are many examples of every possible d electron configuration. What follows is a short description of common geometries and characteristics of each possible d electron count and representative examples.

d0
Commonly tetrahedral however it is possible for d0 complexes to accommodate many electron pairs (bonds/coordination number) since their d orbitals are empty and well away from the 18-electron ceiling. Often colorless due to the lack of d to d transitions. Examples: Titanium tetrachloride, Titanocene dichloride, Schwartz's reagent.

d1
Examples: Molybdenum(V) chloride, Vanadyl acetylacetonate, Vanadocene dichloride, Vanadium tetrachloride.

d2
Examples: Titanocene dicarbonyl.

d3
Examples: Reinecke's salt.

d4
Octahedral high-spin: 4 unpaired electrons, paramagentic, substitutionally labile. Octahedral low-spin: 2 unpaired electrons, paramagentic, substitutionally inert.

d5
Octahedral high-spin: 5 unpaired electrons, paramagentic, substitutionally labile. Octahedral low-spin: 1 unpaired paramagentic, substitutionally inert. electron,

Examples: Potassium ferrioxalate, Vanadium carbonyl.


High-spin [Fe(NO2)6]3 crystal field diagram

d electron count

102

d6
Commonly octahedral complexes in both high spin and low spin. Octahedral high-spin: 4 unpaired electrons, paramagentic, substitutionally labile. Octahedral low-spin: no unpaired electrons, diamagnetic, substitutionally inert.

Examples: Cobalt(III) hexammine chloride, Sodium cobaltinitrite, Molybdenum hexacarbonyl, Ferrocene, Ferroin, Chromium carbonyl.

Low-spin [Fe(NO2)6]3 crystal field diagram

d7
Octahedral high spin: 3 unpaired electrons, paramagentic, substitutionally labile. Octahedral low spin:1 unpaired electron, paramagentic, substitutionally labile. Examples: Cobaltocene.

d8
Complexes which are d8 high-spin are usually 18 electron octahedral while low-spin d8 complexes are generally 16 electron square planar complexes. For first row transition metal complexes such as Ni2+ and Cu+ also form five coordinate 18 electron species which vary from square pyramidal to trigonal bipyramidal. Octahedral high spin: 2 unpaired electrons, paramagentic, substitutionally labile. Square planar low spin: no unpaired electrons, diamagnetic, substitutionally inert. Examples: Cisplatin, Nickelocene, Dichlorobis(ethylenediamine)nickel(II), Iron pentacarbonyl, Zeise's salt, Vaska's complex, Wilkinson's catalyst.

d9
Stable complexes with this electron count is more common for first row (period four) transition metals center than it is for complexes based around second or third row transition metals centers. These include both four coordinate 17 electron species and five coordinate 19 electrons species. Examples: Schweizer's reagent.

d10
Often tetrahedral complexes limited to form 4 additional bonds (8 additional electrons) by the 18-electron ceiling. Often colorless due to the lack of d to d transitions. Examples: Tetrakis(triphenylphosphine)palladium(0), Nickel carbonyl.

d electron count

103

References
[1] Green, M. L. H. (1995-09-20). "A new approach to the formal classification of covalent compounds of the elements". Journal of Organometallic Chemistry 500 (1-2): 127148. doi:10.1016/0022-328X(95)00508-N. ISSN0022-328X. [2] http:/ / www. columbia. edu/ cu/ chemistry/ groups/ parkin/ mlxz. htm [3] Miessler, Gary L.; Donald A. Tarr (1998). Inorganic Chemistry (2nd edition). Upper Saddle River, New Jersey: Pearson Education, Inc. Pearson Prentice Hall. ISBN0138418918. [4] Miessler and Tarr p.38

Hund's rule of maximum multiplicity


Hund's Rule of Maximum Multiplicity is an observational rule which states that a greater total spin state usually makes the resulting atom more stable. Accordingly, it can be taken that if two or more orbitals of equal energy are available, electrons will occupy them singly before filling them in pairs. The rule, discovered by Friedrich Hund in 1925, is of important use in atomic chemistry, spectroscopy, and quantum chemistry. As a result this rule is often abbreviated to Hund's Rule, ignoring Hund's other two rules.

Details
The increased stability of the atom, most commonly manifested in a lower energy state, arises because the high-spin state forces the unpaired electrons to reside in different spatial orbitals. A false, but commonly given, reason for the increased stability of high multiplicity states is that the different occupied spatial orbitals create a larger average distance between electrons, reducing electron-electron repulsion energy. In reality, it has been shown that the actual reason behind the increased stability is a decrease in the screening of electron-nuclear attractions.[1] Total spin state is calculated as the total number of unpaired electrons + 1, or twice the total spin + 1 written as 2S+1. As a result of Hund's rule, constraints are placed on the way atomic orbitals are filled using the Aufbau principle. Before any two electrons occupy an orbital in a subshell, other orbitals in the same subshell must first each contain one electron. Also, the electrons filling a subshell will have parallel spin before the shell starts filling up with the opposite spin electrons (after the first orbital gains a second electron). As a result, when filling up atomic orbitals, the maximum number of unpaired electrons (and hence maximum total spin state) is assured. For example a p4 subshell arranges its electrons as [][][] rather than [][][] or [][][].

Exception
In 2004, researchers reported the synthesis of 5-dehydro-m-xylylene (DMX), the first organic molecule known to violate Hund's rule.[2]

External links
A glossary entry hosted on the web site of the Chemistry Department of Purdue University [3]

References
[1] I.N. Levine, Quantum Chemistry (Prentice-Hall, 4th edn 1991) [ISBN 0205127703], pp. 303-304 [2] Slipchenko, L.; Munsch, T.; Wenthold, P.; Krylov, A. (2004). "5-Dehydro-1,3-quinodimethane: a hydrocarbon with an open-shell doublet ground state". Angewandte Chemie (International ed. in English) 43 (6): 742745. doi:10.1002/anie.200352990. PMID14755709. [3] http:/ / www. chem. purdue. edu/ gchelp/ gloss/ hundsrule. html

Aufbau principle

104

Aufbau principle
The Aufbau principle (from the German Aufbau meaning "building up, construction": also Aufbau rule or building-up principle) is used to determine the electron configuration of an atom, molecule or ion. The principle postulates a hypothetical process in which an atom is "built up" by progressively adding electrons. As they are added, they assume their most stable conditions (electron orbitals) with respect to the nucleus and those electrons already there. According to the principle, electrons fill orbitals starting at the lowest available (possible) energy states before filling higher states (e.g. 1s before 2s). The number of electrons that can occupy each orbital is limited by the Pauli exclusion principle. If multiple orbitals of the same energy are available, Hund's rule states that unoccupied orbitals will be filled before occupied orbitals are reused (by electrons having different spins). A version of the Aufbau principle can also be used to predict the configuration of protons and neutrons in an atomic nucleus.

The Madelung energy ordering rule


The order in which these orbitals are filled is given by the n + l rule (also known as the Madelung rule (after Erwin Madelung), or the Klechkowski rule (after Vsevolod Klechkovsky in some, mostly French- and Russian-speaking, countries), where orbitals with a lower n + l value are filled before those with higher n + l values. In this context, n represents the principal quantum number and l the azimuthal quantum number; the values l = 0, 1, 2, 3 correspond to the s, p, d, and f labels, respectively. The rule is based on the total number of nodes in the atomic orbital, n + l, which is related to the energy.[1] In the case of equal n + l values, the Order in which orbitals are arranged by increasing energy according to the orbital with a lower n value is filled first. The fact Madelung rule. Each diagonal red arrow corresponds to a different value of n + l. that most of the ground state configurations of neutral atoms fill orbitals following this n + l, n pattern was obtained experimentally, by reference to the spectroscopic characteristics of the elements.[2] The Madelung energy ordering rule applies only to neutral atoms in their ground state, and even in that case, there are several elements for which it predicts configurations that differ from those determined experimentally.[3] Copper and chromium are common examples of this property. According to the Madelung rule, the 4s orbital (n + l = 4 + 0 = 4) is occupied before the 3d orbital (n + l = 3 + 2 = 5). The rule then predicts the configuration of 29Cu to be 1s22s22p63s2 3p64s23d9, abbreviated [Ar]4s23d9 where [Ar] denotes the configuration of Ar (the preceding noble gas). However the experimental electronic configuration of the copper atom is [Ar]4s13d10. By filling the 3d orbital, copper can be in a lower energy state. Similarly, chromium takes the electronic configuration of [Ar]4s13d5 instead of [Ar]4s23d4. In this case, chromium has a half-full 3d shell.

Aufbau principle

105

History
The Aufbau principle in the old quantum theory
The principle takes its name from the German, Aufbauprinzip, "building-up principle", rather than being named for a scientist. In fact, it was formulated by Niels Bohr and Wolfgang Pauli in the early 1920s, and states that:

In the old quantum theory, orbits with low angular momentum (s- and p-orbitals) get closer to the nucleus.

The orbitals of lower energy are filled in first with the electrons and only then the orbitals of high energy are filled.

This was an early application of quantum mechanics to the properties of electrons, and explained chemical properties in physical terms. Each added electron is subject to the electric field created by the positive charge of the atomic nucleus and the negative charge of other electrons that are bound to the nucleus. Although in hydrogen there is no energy difference between orbitals with the same principal quantum number n, this is not true for the outer electrons of other atoms. In the old quantum theory prior to quantum mechanics, electrons were supposed to occupy classical elliptical orbita. The orbits with the highest angular momentum are 'circular orbits' outside the inner electrons, but orbits with low angular momentum (s- and p-orbitals) have high orbital eccentricity, so that they get closer to the nucleus and feel on average a less strongly screened nuclear charge.

The n + l energy ordering rule


A periodic table in which each row corresponds to one value of n + l was suggested by Charles Janet in 1927. In 1936, the German physicist Erwin Madelung proposed his empirical rules for the order of filling atomic subshells, based on knowledge of atomic ground states determined by the analysis of atomic spectra, and most English-language sources therefore refer to the Madelung rule. Madelung may have been aware of this pattern as early as 1926.[4] In 1962 the Russian agricultural chemist V.M. Klechkowski proposed the first theoretical explanation for the importance of the sum n + l, based on the statistical Thomas-Fermi model of the atom.[5] Many French-language sources therefore refer to the Klechkowski rule.

Aufbau principle

106

References
[1] Weinhold, Frank; Landis, Clark R. (2005). Valency and bonding: A Natural Bond Orbital Donor-Acceptor Perspective. Cambridge: Cambridge University Press. pp.71516. ISBN0521831288. [2] Scerri, Eric R. (1998). "How Good is the Quantum Mechanical Explanation of the Periodic System?" (http:/ / www. chem. ucla. edu/ dept/ Faculty/ scerri/ pdf/ How_Good_is. pdf). J. Chem. Ed. 75 (11): 138485. Bibcode1998JChEd..75.1384S. doi:10.1021/ed075p1384. . [3] Meek, Terry L.; Allen, Leland C. (2002). "Configuration irregularities: deviations from the Madelung rule and inversion of orbital energy levels" (http:/ / www. sciencedirect. com/ science?_ob=ArticleURL& _udi=B6TFN-46G4S5S-1& _user=961305& _rdoc=1& _fmt=& _orig=search& _sort=d& view=c& _acct=C000049425& _version=1& _urlVersion=0& _userid=961305& md5=cef78ae6aced8ded250c6931a0842063). Chem. Phys. Lett. 362 (56): 36264. Bibcode2002CPL...362..362M. doi:10.1016/S0009-2614(02)00919-3. . [4] Goudsmit, S. A.; Richards, Paul I. (1964). "The Order of Electron Shells in Ionized Atoms" (http:/ / www. pnas. org/ content/ 51/ 4/ 664. full. pdf). Proc. Natl. Acad. Sci. 51 (4): 664671 (with correction on p906). Bibcode1964PNAS...51..664G. doi:10.1073/pnas.51.4.664. . [5] Wong, D. Pan (1979). "Theoretical justification of Madelung's rule" (http:/ / jchemed. chem. wisc. edu/ Journal/ Issues/ 1979/ Nov/ jceSubscriber/ JCE1979p0714. pdf). J. Chem. Ed. 56 (11): 71418. Bibcode1979JChEd..56..714W. doi:10.1021/ed056p714. .

Further reading
Image: Understanding order of shell filling (http://www.iun.edu/~cpanhd/C101webnotes/ modern-atomic-theory/images/energy-levels.jpg) Boeyens, J. C. A.: Chemistry from First Principles. Berlin: Springer Science 2008, ISBN 978-1-4020-8546-8 Scerri, Eric (2006). The Periodic Table: Its Story and Its Significance. New York: Oxford University Press. ISBN0195305736. Ostrovsky, V.N. (2005). "On Recent Discussion Concerning Quantum Justification of the Periodic Table of the Elements" (http://www.springerlink.com/content/p2rqg32684034736/fulltext.pdf). Foundations of Chemistry 7 (3): 23539. doi:10.1007/s10698-005-2141-y. Abstract (http://www.springerlink.com/content/ p2rqg32684034736/fulltext.pdf?page=1). Kitagawara, Y.; Barut, A.O. (1984). "On the dynamical symmetry of the periodic table. II. Modified Demkov-Ostrovsky atomic model" (http://www.iop.org/EJ/article/0022-3700/17/21/013/jbv17i21p4251. pdf). J. Phys. B: At. Mol. Phys. 17 (21): 425159. Bibcode1984JPhB...17.4251K. doi:10.1088/0022-3700/17/21/013.

External links
Electron Configurations, the Aufbau Principle, Degenerate Orbitals, and Hund's Rule (http://chemed.chem. purdue.edu/genchem/topicreview/bp/ch6/quantum.html#aufbau)

Coordination complex

107

Coordination complex
In chemistry, a coordination complex or metal complex, is an atom or ion (usually metallic), bonded to a surrounding array of molecules or anions, that are in turn known as ligands or complexing agents.[1][2] Many metal-containing compounds consist of coordination complexes.[3]

Nomenclature and terminology


Coordination complexes are so pervasive that the structure and reactions are described in many ways, sometimes confusingly. The atom within a ligand that is bonded to the central atom or ion is Cisplatin, PtCl2(NH3)2A platinum atom with four called the donor atom. A typical complex is bound to several ligands donor atoms, which can be the same or different. Polydentate (multiple bonded) ligands consist of several donor atoms, several of which are bound to the central atom or ion. These complexes are called chelate complexes, the formation of such complexes is called chelation, complexation, and coordination. The central atom or ion, together with all ligands comprise the coordination sphere.[4][5] The central atoms or ion and the donor atoms comprise the first coordination sphere. Coordination refers to the "coordinate covalent bonds" (dipolar bonds) between the ligands and the central atom. Originally, a complex implied a reversible association of molecules, atoms, or ions through such weak chemical bonds. As applied to coordination chemistry, this meaning has evolved. Some metal complexes are formed virtually irreversibly and many are bound together by bonds that are quite strong.[6][7]

History
Coordination complexes were known - although not understood in any sense - since the beginning of chemistry, e.g. Prussian blue and copper vitriol. The key breakthrough occurred when Alfred Werner proposed in 1893 that Co(III) bears six ligands in an octahedral geometry. His theory allows one to understand the difference between coordinated and ionic in a compound, for example chloride in the cobalt ammine chlorides and to explain many of the previously inexplicable isomers. In 1914, Werner resolved the first coordination complex, called hexol, into optical isomers, overthrowing the theory that only carbon compounds could possess chirality.
Structure of hexol

Coordination complex

108

Structures
The ions or molecules surrounding the central atom are called ligands. Ligands are generally bound to the central atom by a coordinate covalent bond (donating electrons from a lone electron pair into an empty metal orbital), and are said to be coordinated to the atom. There are also organic ligands such as alkenes whose pi bonds can coordinate to empty metal orbitals. An example is ethene in the complex known as Zeise's salt, K+[PtCl3(C2H4)]-.

Geometry
In coordination chemistry, a structure is first described by its coordination number, the number of ligands attached to the metal (more specifically, the number of donor atoms). Usually one can count the ligands attached, but sometimes even the counting can become ambiguous. Coordination numbers are normally between two and nine, but large numbers of ligands are not uncommon for the lanthanides and actinides. The number of bonds depends on the size, charge, and electron configuration of the metal ion and the ligands. Metal ions may have more than one coordination number. Typically the chemistry of complexes is dominated by interactions between s and p molecular orbitals of the ligands and the d orbitals of the metal ions. The s, p, and d orbitals of the metal can accommodate 18 electrons (see 18-Electron rule). The maximum coordination number for a certain metal is thus related to the electronic configuration of the metal ion (to be more specific, the number of empty orbitals) and to the ratio of the size of the ligands and the metal ion. Large metals and small ligands lead to high coordination numbers, e.g. [Mo(CN)8]4-. Small metals with large ligands lead to low coordination numbers, e.g. Pt[P(CMe3)]2. Due to their large size, lanthanides, actinides, and early transition metals tend to have high coordination numbers. Different ligand structural arrangements result from the coordination number. Most structures follow the points-on-a-sphere pattern (or, as if the central atom were in the middle of a polyhedron where the corners of that shape are the locations of the ligands), where orbital overlap (between ligand and metal orbitals) and ligand-ligand repulsions tend to lead to certain regular geometries. The most observed geometries are listed below, but there are many cases that deviate from a regular geometry, e.g. due to the use of ligands of different types (which results in irregular bond lengths; the coordination atoms do not follow a points-on-a-sphere pattern), due to the size of ligands, or due to electronic effects (see, e.g., Jahn-Teller distortion): Linear for two-coordination Trigonal planar for three-coordination Tetrahedral or square planar for four-coordination Trigonal bipyramidal or square pyramidal for five-coordination Octahedral (orthogonal) or trigonal prismatic for six-coordination Pentagonal bipyramidal for seven-coordination Square antiprismatic for eight-coordination Tri-capped trigonal prismatic (Triaugmented triangular prism) for nine coordination.

Some exceptions and provisions should be noted: The idealized descriptions of 5-, 7-, 8-, and 9- coordination are often indistinct geometrically from alternative structures with slightly different L-M-L (ligand-metal-ligand) angles. The classic example of this is the difference between square pyramidal and trigonal bipyramidal structures. Due to special electronic effects such as (second-order) Jahn-Teller stabilization, certain geometries are stabilized relative to the other possibilities, e.g. for some compounds the trigonal prismatic geometry is stabilized relative to octahedral structures for six-coordination.

Coordination complex

109

Isomerism
The arrangement of the ligands is fixed for a given complex, but in some cases it is mutable by a reaction that forms another stable isomer. There exist many kinds of isomerism in coordination complexes, just as in many other compounds. Stereoisomerism Stereoisomerism occurs with the same bonds in different orientations relative to one another. Stereoisomerism can be further classified into: Cis-trans isomerism and facial-meridional isomerism Cis-trans isomerism occurs in octahedral and square planar complexes (but not tetrahedral). When two ligands are mutually adjacent they are said to be cis, when opposite each other, trans. When three identical ligands occupy one face of an octahedron, the isomer is said to be facial, or fac. In a fac isomer, any two identical ligands are adjacent or cis to each other. If these three ligands and the metal ion are in one plane, the isomer is said to be meridional, or mer. A mer isomer can be considered as a combination of a trans and a cis, since it contains both trans and cis pairs of identical ligands.

cis-[CoCl2(NH3)4]+

trans-[CoCl2(NH3)4]+

fac-[CoCl3(NH3)3

mer-[CoCl3(NH3)3

Optical isomerism Optical isomerism occurs when a molecule is not superposable with its mirror image. It is so called because the two isomers are each optically active, that is, they rotate the plane of polarized light in opposite directions. The symbol (lambda) is used as a prefix to describe the left-handed propeller twist formed by three bidentate ligands, as shown. Likewise, the symbol (delta) is used as a prefix for the right-handed propeller twist.[8]

-[Fe(ox)3]3

-[Fe(ox)3]3

-cis-[CoCl2(en)2]+

-cis-[CoCl2(en)2]+

Coordination complex Structural isomerism Structural isomerism occurs when the bonds are themselves different. There are four types of structural isomerism: ionisation isomerism, solvate or hydrate isomerism, linkage isomerism and coordination isomerism. 1. Ionisation isomerism - the isomers give different ions in solution although they have the same composition. This type of isomerism occurs when the counter ion of the complex is also a potential ligand. For example pentaaminebromidocobalt(III)sulphate [Co(NH3)5Br]SO4 is red violet and in solution gives a precipitate with barium chloride, confirming the presence of sulphate ion, while pentaaminesulphatecobalt(III)bromide {Co(NH3)5SO4]Br is red and tests negative for sulphate ion in solution, but instead gives a precipitate of AgBr with silver nitrate. 2. Solvate or hydrate isomerism - the isomers have the same composition but differ with respect to the number of solvent ligand molecules as well as the counter ion in the crystal lattice. For example [Cr(H2O)6]Cl3 is violet colored, [Cr(H2O)5Cl]Cl2.H2O is blue-green, and [Cr(H2O)4Cl2]Cl.2H2O is dark green 3. Linkage isomerism occurs with ambidentate ligands that can bind in more than one place. For example, NO2 is an ambidentate ligand: It can bind to a metal at either the N atom or an O atom. 4. Coordination isomerism - this occurs when both positive and negative ions of a salt are complex ions and the two isomers differ in the distribution of ligands between the cation and the anion. For example [Co(NH3)6][Cr(CN)6] and [Cr(NH3)6][Co(CN)6]

110

Electronic properties
Many of the properties of metal complexes are dictated by their electronic structures. The electronic structure can be described by a relatively ionic model that ascribes formal charges to the metals and ligands. This approach is the essence of crystal field theory (CFT). Crystal field theory, introduced by Hans Bethe in 1929, gives a quantum mechanically based attempt at understanding complexes. But crystal field theory treats all interactions in a complex as ionic and assumes that the ligands can be approximated by negative point charges. More sophisticated models embrace covalency, and this approach is described by ligand field theory (LFT) and Molecular orbital theory (MO). Ligand field theory, introduced in 1935 and built from molecular orbital theory, can handle a broader range of complexes and can explain complexes in which the interactions are covalent. The chemical applications of group theory can aid in the understanding of crystal or ligand field theory, by allowing simple, symmetry based solutions to the formal equations. Chemists tend to employ the simplest model required to predict the properties of interest; for this reason, CFT has been a favorite for the discussions when possible. MO and LF theories are more complicated, but provide a more realistic perspective. The electronic configuration of the complexes gives them some important properties:

Coordination complex

111

Color
Metal complexes often have spectacular colors caused by electronic transitions by the absorption of light. For this reason they are often applied as pigments. Most transitions that are related to colored metal complexes are either d-d transitions or charge transfer bands. In a d-d transition, an electron in a d orbital on the metal is excited by a photon to another d orbital of higher energy. A charge transfer band entails promotion of an electron from a metal-based orbital into an empty ligand-based orbital (Metal-to-Ligand Charge Transfer or MLCT). The converse also occurs: excitation of an electron in a Synthesis of copper(II)-tetraphenylporphyrin, a ligand-based orbital into an empty metal-based orbital (Ligand to metal complex, from tetraphenylporphyrin and Metal Charge Transfer or LMCT). These phenomena can be observed copper(II) acetate monohydrate. with the aid of electronic spectroscopy; also known as UV-Vis.[9] For simple compounds with high symmetry, the d-d transitions can be assigned using Tanabe-Sugano diagrams. These assignments are gaining increased support with computational chemistry.

Colours of Various Example Coordination Complexes


FeII Hydrated Ion [Fe(H2O)6]2+ Pale green Soln FeIII [Fe(H2O)6]3+ Yellow/brown Soln CoII [Co(H2O)6]2+ Pink Soln CuII [Cu(H2O)6]2+ Blue Soln AlIII [Al(H2O)6]3+ Colourless Soln CrIII [Cr(H2O)6]3+ Green Soln

OH, dilute

[Fe(H2O)4(OH)2] [Fe(H2O)3(OH)3] [Co(H2O)4(OH)2] [Cu(H2O)4(OH)2] Dark green Brown Blue/green Blue Ppt Ppt Ppt Ppt

[Al(H2O)3(OH)3] [Cr(H2O)3(OH)3] White Green Ppt Ppt [Al(OH)4] Colourless Soln [Cr(OH)6]3 Green Soln

OH, concentrated [Fe(H2O)4(OH)2] [Fe(H2O)3(OH)3] [Co(H2O)4(OH)2] [Cu(H2O)4(OH)2] Dark green Brown Blue/green Blue Ppt Ppt Ppt Ppt NH3, dilute [Fe(H2O)4(OH)2] [Fe(H2O)3(OH)3] [Co(H2O)4(OH)2] [Cu(H2O)4(OH)2] Dark green Brown Blue/green Blue Ppt Ppt Ppt Ppt

[Al(H2O)3(OH)3] [Cr(H2O)3(OH)3] White Green Ppt Ppt

NH3, concentrated [Fe(H2O)4(OH)2] [Fe(H2O)3(OH)3] [Co(NH ) ]2+ 3 6 Dark green Brown Straw coloured Ppt Ppt Soln CO32 FeCO3 Dark green Ppt [Fe(H2O)3(OH)3] CoCO3 Brown Pink Ppt + bubbles Ppt

[Cu(NH3)4(H2O)2]2+ [Al(H2O)3(OH)3] [Cr(NH3)6]3+ White Deep blue Green Ppt Soln Soln CuCO3 Blue/green Ppt

Magnetism
Metal complexes that have unpaired electrons are magnetic. Considering only monometallic complexes, unpaired electrons arise because the complex has an odd number of electrons or because electron pairing is destabilized. Thus, monomeric Ti(III) species have one "d-electron" and must be (para)magnetic, regardless of the geometry or the nature of the ligands. Ti(II), with two d-electrons, forms some complexes that have two unpaired electrons and others with none. This effect is illustrated by the compounds TiX2[(CH3)2PCH2CH2P(CH3)2]2: when X = Cl, the complex is paramagnetic (high-spin configuration), whereas when X=CH3, it is diamagnetic (low-spin configuration). It is important to realize that ligands provide an important means of adjusting the ground state

Coordination complex properties. In bi- and polymetallic complexes, in which the individual centers have an odd number of electrons or that are high-spin, the situation is more complicated. If there is interaction (either direct or through ligand) between the two (or more) metal centers, the electrons may couple (antiferromagnetic coupling, resulting in a diamagnetic compound), or they may enhance each other (ferromagnetic coupling). When there is no interaction, the two (or more) individual metal centers behave as if in two separate molecules.

112

Reactivity
Complexes show a variety of possible reactivities: Electron transfers A common reaction between coordination complexes involving ligands are inner and outer sphere electron transfers. They are two different mechanisms of electron transfer redox reactions, largely defined by the late Henry Taube. In an inner sphere reaction, a ligand with two lone electron pairs acts as a bridging ligand, a ligand to which both coordination centres can bond. Through this, electrons are transferred from one centre to another. (Degenerate) ligand exchange One important indicator of reactivity is the rate of degenerate exchange of ligands. For example, the rate of interchange of coordinate water in [M(H2O)6]n+ complexes varies over 20 orders of magnitude. Complexes where the ligands are released and rebound rapidly are classified as labile. Such labile complexes can be quite stable thermodynamically. Typical labile metal complexes either have low-charge (Na+), electrons in d-orbitals that are antibonding with respect to the ligands (Zn2+), or lack covalency (Ln3+, where Ln is any lanthanide). The lability of a metal complex also depends on the high-spin vs. low-spin configurations when such is possible. Thus, high-spin Fe(II) and Co(III) form labile complexes, whereas low-spin analogues are inert. Cr(III) can exist only in the low-spin state (quartet), which is inert because of its high formal oxidation state, absence of electrons in orbitals that are M-L antibonding, plus some "ligand field stabilization" associated with the d3 configuration. Associative processes Complexes that have unfilled or half-filled orbitals often show the capability to react with substrates. Most substrates have a singlet ground-state; that is, they have lone electron pairs (e.g., water, amines, ethers), so these substrates need an empty orbital to be able to react with a metal centre. Some substrates (e.g., molecular oxygen) have a triplet ground state, which results that metals with half-filled orbitals have a tendency to react with such substrates (it must be said that the dioxygen-molecule also has lone pairs, so it is also capable to react as a 'normal' Lewis base). If the ligands around the metal are carefully chosen, the metal can aid in (stoichiometric or catalytic) transformations of molecules or be used as a sensor.

Classification
Metal complexes, also known as coordination compounds, include all metal compounds, aside from metal vapors, plasmas, and alloys. The study of "coordination chemistry" is the study of "inorganic chemistry" of all alkali and alkaline earth metals, transition metals, lanthanides, actinides, and metalloids. Thus, coordination chemistry is the chemistry of the majority of the periodic table. Metals and metal ions exist, in the condensed phases at least, only surrounded by ligands. The areas of coordination chemistry can be classified according to the nature of the ligands, in broad terms: Classical (or "Werner Complexes"): Ligands in classical coordination chemistry bind to metals, almost exclusively, via their "lone pairs" of electrons residing on the main group atoms of the ligand. Typical ligands are

Coordination complex H2O, NH3, Cl, CN, en Examples: [Co(EDTA)], [Co(NH3)6]Cl3, [Fe(C2O4)3]K3 Organometallic Chemistry: Ligands are organic (alkenes, alkynes, alkyls) as well as "organic-like" ligands such as phosphines, hydride, and CO. Example: (C5H5)Fe(CO)2CH3 Bioinorganic Chemistry: Ligands are those provided by nature, especially including the side chains of amino acids, and many cofactors such as porphyrins. Example: hemoglobin Many natural ligands are "classical" especially including water. Cluster Chemistry: Ligands are all of the above also include other metals as ligands. Example Ru3(CO)12 In some cases there are combinations of different fields: Example: [Fe4S4(Scysteinyl)4]2, in which a cluster is embedded in a biologically active species. Mineralogy, materials science, and solid state chemistry - as they apply to metal ions - are subsets of coordination chemistry in the sense that the metals are surrounded by ligands. In many cases these ligands are oxides or sulfides, but the metals are coordinated nonetheless, and the principles and guidelines discussed below apply. In hydrates, at least some of the ligands are water molecules. It is true that the focus of mineralogy, materials science, and solid state chemistry differs from the usual focus of coordination or inorganic chemistry. The former are concerned primarily with polymeric structures, properties arising from a collective effects of many highly interconnected metals. In contrast, coordination chemistry focuses on reactivity and properties of complexes containing individual metal atoms or small ensembles of metal atoms.

113

Older classifications of isomerism


Traditional classifications of the kinds of isomer have become archaic with the advent of modern structural chemistry. In the older literature, one encounters: Ionisation isomerism describes the possible isomers arising from the exchange between the outer sphere and inner sphere. This classification relies on an archaic classification of the inner and outer sphere. In this classification, the "outer sphere ligands," when ions in solution, may be switched with "inner sphere ligands" to produce an isomer. Solvation isomerism occurs when an inner sphere ligand is replaced by a solvent molecule. This classification is obsolete because it considers solvents as being distinct from other ligands. Some of the problems are discussed under water of crystallization.

Naming complexes
The basic procedure for naming a complex: 1. When naming a complex ion, the ligands are named before the metal ion. 2. Write the names of the ligands in alphabetical order. (Numerical prefixes do not affect the order.) Multiple occurring monodentate ligands receive a prefix according to the number of occurrences: di-, tri-, tetra-, penta-, or hexa. Polydentate ligands (e.g., ethylenediamine, oxalate) receive bis-, tris-, tetrakis-, etc. Anions end in ido. This replaces the final 'e' when the anion ends with '-ate', e.g. sulfate becomes sulfato. It replaces 'ide': cyanide becomes cyanido. Neutral ligands are given their usual name, with some exceptions: NH3 becomes ammine; H2O becomes aqua or aquo; CO becomes carbonyl; NO becomes nitrosyl.

Coordination complex 3. Write the name of the central atom/ion. If the complex is an anion, the central atom's name will end in -ate, and its Latin name will be used if available (except for mercury). 4. If the central atom's oxidation state needs to be specified (when it is one of several possible, or zero), write it as a Roman numeral (or 0) in parentheses. 5. Name cation then anion as separate words (if applicable, as in last example) Examples: [NiCl4]2 tetrachloridonickelate(II) ion [CuNH3Cl5]3 amminepentachloridocuprate(II) ion [Cd(en)2(CN)2] dicyanidobis(ethylenediamine)cadmium(II) [Co(NH3)5Cl]SO4 pentaamminechloridocobalt(III) sulfate The coordination number of ligands attached to more than one metal (bridging ligands) is indicated by a subscript to the Greek symbol placed before the ligand name. Thus the dimer of aluminium trichloride is described by Al2Cl4(2-Cl)2.

114

Application of coordination compounds


1. They are used in photography, i.e., AgBr forms a soluble complex with sodium thiosulfate in photography. 2. 3. 4. 5. K[Ag(CN)2] is used for electroplating of silver, and K[Au(CN)2]is used for gold plating. Some ligands oxidise Co2+ to Co3+ ion. Ethylenediaminetetraacetic acid (EDTA) is used for estimation of Ca2+ and Mg2+in hard water. Silver and gold are extracted by treating zinc with their cyanide complexes.

External links
Naming Coordination Compounds [10] Transition Metal Complex Colors [11]

References
[1] Nic, M.; Jirat, J.; Kosata, B., eds. (2006). "complex" (http:/ / goldbook. iupac. org/ C01203. html). IUPAC Compendium of Chemical Terminology (Online ed.). doi:10.1351/goldbook.C01203. ISBN0-9678550-9-8. . [2] Nic, M.; Jirat, J.; Kosata, B., eds. (2006). "coordination entity" (http:/ / goldbook. iupac. org/ C01330. html). IUPAC Compendium of Chemical Terminology (Online ed.). doi:10.1351/goldbook.C01330. ISBN0-9678550-9-8. . [3] Greenwood, N. N.; Earnshaw, A. (1997). Chemistry of the Elements (2nd ed.). Butterworth-Heinemann. ISBN0080379419. [4] chemistry-dictionary.com - Definition of coordination sphere (http:/ / www. chemistry-dictionary. com/ definition/ coordination+ sphere. php) [5] What Is A Coordination Compound? (http:/ / www. chem. purdue. edu/ gchelp/ cchem/ whatis. html) [6] Cotton, Frank Albert; Geoffrey Wilkinson, Carlos A. Murillo (1999). Advanced Inorganic Chemistry. pp.1355. ISBN0471199575, 9780471199571. [7] Miessler, Gary L.; Donald Arthur Tarr (1999). Inorganic Chemistry. pp.642. ISBN0138418918, 9780138418915. [8] Miessler, Gary L.; Donald Arthur Tarr (1999). "9". Inorganic Chemistry. pp.315, 316. ISBN0138418918, 9780138418915. [9] Harris, D., Bertolucci, M., Symmetry and Spectroscopy. 1989 New York, Dover Publications [10] http:/ / www. chemistry. wustl. edu/ ~edudev/ LabTutorials/ naming_coord_comp. html [11] http:/ / www. wou. edu/ las/ physci/ ch462/ tmcolors. htm De Vito, D.; Weber, J. ; Merbach, A. E. Calculated Volume and Energy Profiles for Water Exchange on t2g 6

Rhodium(III) and Iridium(III) Hexaaquaions: Conclusive Evidence for an Ia Mechanism Inorganic Chemistry, 2005, Volume 43, pages 858-863. Zumdahl, Steven S. Chemical Principles, Fifth Edition. New York: Houghton Mifflin, 2005. 943-946, 957. Harris, D., Bertolucci, M., Symmetry and Spectroscopy. 1989 New York, Dover Publications

Diamagnetism

115

Diamagnetism
Diamagnetism is the property of an object or material which causes it to create a magnetic field in opposition to an externally applied magnetic field. Diamagnetism is believed to be due to quantum mechanics (and is understood in terms of Landau levels[1]) and occurs because the external field alters the orbital velocity of electrons around their nuclei, thus changing the magnetic dipole moment. According to Lenz's law, the field of these electrons will oppose the magnetic field changes provided by the applied field. In most materials diamagnetism is a weak effect, but in a superconductor a strong quantum effect repels the magnetic field entirely, apart from a thin layer at the surface. Diamagnets are materials with a magnetic permeability less than (a relative permeability less than1).

History

Levitating pyrolytic carbon

In 1778, Sebald Justinus Brugmans was the first individual to observe that bismuth and antimony were repelled by magnetic fields. However, the term diamagnetism was coined by Michael Faraday in September 1845, when he realized that every material responded (in either a diamagnetic or paramagnetic way) to an applied magnetic field.

Diamagnetic materials
Notable diamagnetic materials[2]
Material Superconductor Pyrolytic carbon Bismuth Mercury Silver v (105) 105 40.0 16.6 2.9 2.6

Carbon (diamond) 2.1 Lead 1.8

Carbon (graphite) 1.6 Copper Water 1.0 0.91

Diamagnetism is a very general phenomenon, because all electrons, including the electrons of an atom, will always make a weak contribution to the material's response. However, for materials that show some other form of magnetism (such as ferromagnetism or paramagnetism), the diamagnetism is completely overpowered. Substances that mostly display diamagnetic behaviour are termed diamagnetic materials, or diamagnets. Materials that are said

Diamagnetism to be diamagnetic are those that are usually considered by non-physicists to be non-magnetic, and include water, wood, most organic compounds such as petroleum and some plastics, and many metals including copper, particularly the heavy ones with many core electrons, such as mercury, gold and bismuth. The magnetic susceptibility of various molecular fragments are called Pascal's constants. Diamagnetic materials have a relative magnetic permeability that is less than or equal to 1, and therefore a magnetic susceptibility which is less than 0 since susceptibility is defined as v=v1. This means that diamagnetic materials are repelled by magnetic fields. However, since diamagnetism is such a weak property its effects are not observable in everyday life. For example, the magnetic susceptibility of diamagnets such as water is v = 9.05106. The most strongly diamagnetic material is bismuth, v = 1.66104, although pyrolytic carbon may have a susceptibility of v = 4.00104 in one plane. Nevertheless, these values are orders of magnitudes smaller than the magnetism exhibited by paramagnets and ferromagnets. Note that because v is derived from the ratio of the internal magnetic field to the applied field, it is a dimensionless value. All conductors exhibit an effective diamagnetism when they experience a changing magnetic field. The Lorentz force on electrons causes them to circulate around forming eddy currents. The eddy currents then produce an induced magnetic field which opposes the applied field, resisting the conductor's motion. Superconductors may be considered to be perfect diamagnets (v = 1), since they expel all fields (except in a thin surface layer) due to the Meissner effect. However this effect is not due to eddy currents, as in ordinary diamagnetic materials (see the article on superconductivity).

116

Demonstrations of diamagnetism
Curving water surfaces
A superconductor acts as an essentially perfect diamagnetic material when placed in a magnetic field and it excludes the field, and the flux lines avoid the region

If a powerful magnet (such as a supermagnet) is covered with a layer of water (that is thin compared to the diameter of the magnet) then the field of the magnet significantly repels the water. This causes a slight dimple in the water's surface that may be seen by its reflection.[3][4]

Diamagnetic levitation

Diamagnetism

117 Diamagnets may be levitated in stable equilibrium in a magnetic field, with no power consumption. Earnshaw's theorem seems to preclude the possibility of static magnetic levitation. However, Earnshaw's theorem only applies to objects with positive moments, such as ferromagnets (which have a permanent positive moment) and paramagnets (which induce a positive moment). These are attracted to field maxima, which do not exist in free space. Diamagnets (which induce a negative moment) are attracted to field minima, and there can be a field minimum in free space. A thin slice of pyrolytic graphite, which is an unusually strong diamagnetic material, can be stably floated in a magnetic field, such as that from rare earth permanent magnets. This can be done with all components at room temperature, making a visually effective demonstration of diamagnetism.

The Radboud University Nijmegen, the Netherlands, has conducted experiments where water and other substances were successfully levitated. Most spectacularly, a live frog (see figure) was levitated.[6] In September 2009, NASA's Jet Propulsion Laboratory in Pasadena, California announced they had successfully levitated mice using a superconducting magnet,[7] an important step forward since mice are closer biologically to humans than frogs.[8] They hope to perform experiments regarding the effects of microgravity on bone and muscle mass. Recent experiments studying the growth of protein crystals has led to a technique using powerful magnets to allow growth in ways that counteract Earth's gravity.[9] A simple homemade device for demonstration can be constructed out of bismuth plates and a few permanent magnets that will levitate a permanent magnet.[10]

A live frog levitates inside a 32 mm diameter vertical bore of a Bitter solenoid in a magnetic field of about 16 teslas at the Nijmegen High Field Magnet [5] Laboratory.

Theory of diamagnetism
The Bohrvan Leeuwen theorem proves that there cannot be any diamagnetism or paramagnetism in a purely classical system. Yet the classical theory for Langevin diamagnetism gives the same prediction as the quantum theory.[11] The classical theory is given below.

Langevin diamagnetism
The Langevin theory of diamagnetism applies to materials containing atoms with closed shells (see dielectrics). A field with intensity B, applied to an electron with charge e and mass m, gives rise to Larmor precession with frequency = eB / 2m. The number of revolutions per unit time is / 2, so the current for an atom with Z electrons is (in SI units)[11]

The magnetic moment of a current loop is equal to the current times the area of the loop. Suppose the field is aligned with the z axis. The average loop area can be given as , where is the mean square distance of the electrons perpendicular to the z axis. The magnetic moment is therefore

Diamagnetism If the distribution of charge is spherically symmetric, we can suppose that the distribution of x,y,z coordinates are independent and identically distributed. Then , where is the mean square distance of the electrons from the nucleus. Therefore volume, the diamagnetic susceptibility is . If is the number of atoms per unit

118

Diamagnetism in metals
The Langevin theory does not apply to metals because they have non-localized electrons. The theory for the diamagnetism of a free electron gas is called Landau diamagnetism, and instead considers the weak counter-acting field that forms when their trajectories are curved due to the Lorentz force. Landau diamagnetism, however, should be contrasted with Pauli paramagnetism, an effect associated with the polarization of delocalized electrons' spins.[12]

References
[1] http:/ / physics. ucsc. edu/ ~peter/ 231/ magnetic_field/ node5. html [2] Nave, Carl L.. "Magnetic Properties of Solids" (http:/ / hyperphysics. phy-astr. gsu. edu/ Hbase/ tables/ magprop. html). Hyper Physics. . Retrieved 2008-11-09. [3] Beatty, Bill (2005). "Neodymium supermagnets: Some demonstrationsDiamagnetic water" (http:/ / amasci. com/ amateur/ neodymium. html#water). Science Hobbyist. . Retrieved September 2011. [4] Quit007 (2011). "Diamagnetism Gallery" (http:/ / quit007. deviantart. com/ gallery/ 23787987). DeviantART. . Retrieved September 2011. [5] "The Frog That Learned to Fly" (http:/ / www. ru. nl/ hfml/ research/ levitation/ diamagnetic/ ). High Field Laboratory. Radboud University Nijmegen. 2011. . Retrieved September 2011. [6] "The Real Levitation" (http:/ / www. ru. nl/ hfml/ research/ levitation/ diamagnetic/ ). High Field Laboratory. Radboud University Nijmegen. 2011. . Retrieved September 2011. [7] Liu, Yuanming; Zhu, Da-Ming; Strayer, Donald M.; Israelsson, Ulf E. (2010). "Magnetic levitation of large water droplets and mice". Advances in Space Research 45 (1): 208213. Bibcode2010AdSpR..45..208L. doi:10.1016/j.asr.2009.08.033. [8] Choi, Charles Q. (09-09-2009). "Mice levitated in lab" (http:/ / www. livescience. com/ animals/ 090909-mouse-levitation. html). Live Science. . Retrieved September 2011. [9] Kleiner, Kurt (08-10-2007). "Magnetic gravity trick grows perfect crystals" (http:/ / www. newscientist. com/ article/ dn12467-magnetic-gravity-trick-grows-perfect-crystals. html). New Scientist. . Retrieved September 2011. [10] "Fun with diamagnetic levitation" (http:/ / web. archive. org/ web/ 20080212011654/ http:/ / www. fieldlines. com/ other/ diamag1. html). ForceField. 02-12-2008. . Retrieved September 2011. [11] Kittel, Charles (1986). Introduction to Solid State Physics (6th ed.). John Wiley & Sons. pp.299302. ISBN0-471-87474-4. [12] Chang, M. C.. "Diamagnetism and paramagnetism" (http:/ / phy. ntnu. edu. tw/ ~changmc/ Teach/ SS/ SS_note/ chap11. pdf). NTNU lecture notes. . Retrieved 2011-02-24.

External links
Video of a museum-style magnetic elevation train model which makes use of diamagnetism (http://www. youtube.com/watch?v=8tFsrGRwOOM) Videos of frogs and other diamagnets levitated in a strong magnetic field (http://www.ru.nl/hfml/research/ levitation/diamagnetic/) Video of levitating pyrolytic graphite (http://www.grand-illusions.com/images/articles/toyshop/ diamagnetic_levitation_2/diamagnetic_levitation_2.wmv) Video of Meissner-Ochsenfeld effect involving liquid nitrogen (http://www.science.tv/watch/ e257e44aa9d5bade97ba/liquid-nitrogen-and-superconductor) Video of a piece of neodymium magnet levitating between blocks of bismuth. (http://netti.nic.fi/~054028/ images/LevitorMK1.0-1.mpg) Website about this device, with images (in Finnish). (http://netti.nic.fi/~054028/)

Paramagnetism

119

Paramagnetism
Paramagnetism is a form of magnetism whereby the paramagnetic material is only attracted when in the presence of an externally applied magnetic field. In contrast with this behavior, diamagnetic materials are repelled by magnetic fields.[1] Paramagnetic materials have a relative magnetic permeability greater or equal to unity (i.e., a positive magnetic susceptibility) and hence are attracted to magnetic fields. The magnetic moment induced by the applied field is linear in the field strength and rather weak. It typically requires a sensitive analytical balance to detect the effect and modern measurements on paramagnetic materials are often conducted with a SQUID magnetometer. Paramagnetic materials have a small, positive susceptibility to magnetic fields. These materials are slightly attracted by a magnetic field and the material does not retain the magnetic properties when the external field is removed. Paramagnetic properties are due to the presence of some unpaired electrons, and from the realignment of the electron paths caused by the external magnetic field. Paramagnetic materials include magnesium, molybdenum, lithium, and tantalum. Unlike ferromagnets, paramagnets do not retain any magnetization in the absence of an externally applied magnetic field, because thermal A trickle of liquid oxygen is deflected by a magnetic field, illustrating its paramagnetic motion randomizes the spin orientations. Some paramagnetic materials property retain spin disorder at absolute zero, meaning they are paramagnetic in the ground state. Thus the total magnetization drops to zero when the applied field is removed. Even in the presence of the field there is only a small induced magnetization because only a small fraction of the spins will be oriented by the field. This fraction is proportional to the field strength and this explains the linear dependency. The attraction experienced by ferromagnetic materials is non-linear and much stronger, so that it is easily observed, for instance, by the attraction between a refrigerator magnet and the iron of the refrigerator itself.

Simple illustration of a paramagnetic probe made up from miniature magnets.

Relation to electron spins


Constituent atoms or molecules of paramagnetic materials have permanent magnetic moments (dipoles), even in the absence of an applied field. The permanent moment generally is due to the spin of unpaired electrons in atomic or molecular electron orbitals (see Magnetic moment). In pure paramagnetism, the dipoles do not interact with one another and are randomly oriented in the absence of an external field due to thermal agitation, resulting in zero net magnetic moment. When a magnetic field is applied, the dipoles will tend to align with the applied field, resulting in a net magnetic moment in the direction of the applied field. In the classical description, this alignment can be understood to occur due to a torque being provided on the magnetic moments by an applied field, which tries to align the dipoles parallel to the applied field. However, the true origins of the alignment can only be understood via the quantum-mechanical properties of spin and angular momentum. If there is sufficient energy exchange between neighbouring dipoles they will interact, and may spontaneously align or anti-align and form magnetic domains, resulting in ferromagnetism (permanent magnets) or antiferromagnetism, respectively. Paramagnetic behavior can also be observed in ferromagnetic materials that are above their Curie temperature, and in antiferromagnets above their Nel temperature. At these temperatures the available thermal energy simply overcomes the interaction energy between the spins.

Paramagnetism In general paramagnetic effects are quite small: the magnetic susceptibility is of the order of 103 to 105 for most paramagnets, but may be as high as 101 for synthetic paramagnets such as ferrofluids.

120

Delocalization Selected Pauli-paramagnetic metals[2]


Material Tungsten Cesium Magnetic susceptibility (105) 6.8 5.1

Aluminium 2.2 Lithium 1.4

Magnesium 1.2 Sodium 0.72

In many metallic materials the electrons are itinerant, i.e. they travel through the solid more or less as an electron gas. This behavior results from strong interactions (good orbital overlap in a chemist's vocabulary) between the wave functions of neighboring atoms in the extended lattice structure. The wave functions of the valence electrons thus form a band with equal numbers of spins up and down. When exposed to an external field only those electrons close to the Fermi level will respond and a small surplus of one type of spins will result. This effect is a weak form of paramagnetism known as Pauli-paramagnetism. The effect always competes with a diamagnetic response of opposite sign due to all the core electrons of the atoms. Stronger forms of magnetism usually require localized rather than itinerant electrons. However in some cases a bandstructure can result in which there are two delocalized sub-bands with states of opposite spins that have different energies. If one subband is preferentially filled over the other, one can have itinerant ferromagnetic order. This situation usually only occurs in relatively narrow (d-)bands, which are poorly delocalized. s and p electrons Generally, strong delocalization in a solid due to large overlap with neighboring wave functions tends to lead to pairing of spins (quenching) and thus weak magnetism. This is why s- and p-type metals are typically either Pauli-paramagnetic or as in the case of gold even diamagnetic. In the latter case the diamagnetic contribution from the closed shell inner electrons simply wins from the weak paramagnetic term of the almost free electrons. d and f electrons Stronger magnetic effects are typically only observed when d- or f-electrons are involved. Particularly the latter are usually strongly localized. Moreover the size of the magnetic moment on a lanthanide atom can be quite large as it can carry up to 7 unpaired electrons in the case of gadolinium(III) (hence its use in MRI). This high magnetic moments associated with lanthanides is one reason why superstrong magnets are typically based on elements like neodymium or samarium. Molecular localization Of course the above picture is a generalization as it pertains to materials with an extended lattice rather than a molecular structure. Molecular structure can also lead to localization of electrons. Although there are usually energetic reasons why a molecular structure results such that it does not exhibit partly filled orbitals (i.e. unpaired spins), some non-closed shell moieties do occur in nature. Molecular oxygen is a good example. Even in the frozen solid it contains di-radical molecules resulting in paramagnetic behavior. The unpaired spins reside in orbitals

Paramagnetism derived from oxygen p wave functions, but the overlap is limited to the one neighbor in the O2 molecules. The distances to other oxygen atoms in the lattice remain too large to lead to delocalization and the magnetic moments remain unpaired.

121

Curie's law
For low levels of magnetization, the magnetization of paramagnets follows what is known as Curie's law, at least approximately. This law indicates that the susceptibility of paramagnetic materials is inversely proportional to their temperature, i.e. that materials become more magnetic at lower temperatures. The mathematical expression is:

where: M is the resulting magnetization is the magnetic susceptibility H is the auxiliary magnetic field, measured in amperes/meter T is absolute temperature, measured in kelvins C is a material-specific Curie constant Curie's law is valid under the commonly encountered conditions of low magnetization (BH kBT), but does not apply in the high-field/low-temperature regime where saturation of magnetization occurs (BH kBT) and magnetic dipoles are all aligned with the applied field. When the dipoles are aligned, increasing the external field will not increase the total magnetization since there can be no further alignment. For a paramagnetic ion with noninteracting magnetic moments with angular momentum J, the Curie constant is related the individual ions' magnetic moments, . The parameter eff is interpreted as the effective magnetic moment per paramagnetic ion. If one uses a classical treatment with molecular magnetic moments represented as discrete magnetic dipoles, , a Curie Law expression of the same form will emerge with appearing in place of eff. Click "show" to see a derivation of this law: Curie's Law can be derived by considering a substance with noninteracting magnetic moments with angular momentum J. If orbital contributions to the magnetic moment are negligible (a common case), then in what follows J = S. If we apply a magnetic field along what we choose to call the z-axis, the energy levels of each paramagnetic center will experience Zeeman splitting of its energy levels, each with a z-component labeled by MJ (or just MS for the spin-only magnetic case). Applying semiclassical Boltzmann statistics, the molar magnetization of such a substance is

Where

is the z-component of the magnetic moment for each Zeeman level, so

B is

called the Bohr Magneton and gJ is the Land g-factor, which reduces to the free-electron g-factor, gS when J = S. (in this treatment, we assume that the x- and y-components of the magnetization, averaged over all molecules, cancel out because the field applied along the z-axis leave them randomly oriented.) The energy of each Zeeman level is . For temperatures over a few K, , and we can apply the approximation :

Paramagnetism

122

which yields: . The molar , and the molar susceptibility is given by . When orbital angular momentum contributions to the magnetic moment are small, as occurs for most organic radicals or for octahedral transition metal complexes with d3 or high-spin d5 configurations, the effective magnetic moment takes the form (ge = 2.0023... 2), , where n is the number of unpaired electrons. In other transition metal complexes this yields a useful, if somewhat cruder, estimate. bulk magnetization is then

Examples of paramagnets
Materials that are called 'paramagnets' are most often those that exhibit, at least over an appreciable temperature range, magnetic susceptibilities that adhere to the Curie or CurieWeiss laws. In principle any system that contains atoms, ions, or molecules with unpaired spins can be called a paramagnet, but the interactions between them need to be carefully considered.

Systems with minimal interactions


The narrowest definition would be: a system with unpaired spins that do not interact with each other. In this narrowest sense, the only pure paramagnet is a dilute gas of monatomic hydrogen atoms. Each atom has one non-interacting unpaired electron. Of course, the latter could be said about a gas of lithium atoms but these already possess two paired core electrons that produce a diamagnetic response of opposite sign. Strictly speaking Li is a mixed system therefore, although admittedly the diamagnetic component is weak and often neglected. In the case of heavier elements the diamagnetic contribution becomes more important and in the case of metallic gold it dominates the properties. Of course, the element hydrogen is virtually never called 'paramagnetic' because the monatomic gas is stable only at extremely high temperature; H atoms combine to form molecular H2 and in so doing, the magnetic moments are lost (quenched), because the spins pair. Hydrogen is therefore diamagnetic and the same holds true for most elements. Although the electronic configuration of the individual atoms (and ions) of most elements contain unpaired spins, it is not correct to call these elements 'paramagnets' because at ambient temperature quenching is very much the rule rather than the exception. However, the quenching tendency is weakest for f-electrons because f (especially 4f) orbitals are radially contracted and they overlap only weakly with orbitals on adjacent atoms. Consequently, the lanthanide elements with incompletely filled 4f-orbitals are paramagnetic or magnetically ordered.[3]

Paramagnetism

123

eff values for typical d3 and d5 transition metal complexes.[4]


Material [Cr(NH3)6]Br3 K3[Cr(CN)6] K3[MoCl6] K4[V(CN)6] [Mn(NH3)6]Cl2 eff/B 3.77 3.87 3.79 3.78 5.92

(NH4)2[Mn(SO4)2]6H2O 5.92 NH4[Fe(SO4)2]12H2O 5.89

Thus, condensed phase paramagnets are only possible if the interactions of the spins that lead either to quenching or to ordering are kept at bay by structural isolation of the magnetic centers. There are two classes of materials for which this holds: Molecular materials with a (isolated) paramagnetic center. Good examples are coordination complexes of d- or f-metals or proteins with such centers, e.g. myoglobin. In such materials the organic part of the molecule acts as an envelope shielding the spins from their neighbors. Small molecules can be stable in radical form, oxygen O2 is a good example. Such systems are quite rare because they tend to be rather reactive. Dilute systems. Dissolving a paramagnetic species in a diamagnetic lattice at small concentrations, e.g. Nd3+ in CaCl2 will separate the neodymium ions at large enough distances that they do not interact. Such systems are of prime importance for what can be considered the most sensitive method to study paramagnetic systems: EPR.

Systems with interactions


As stated above many materials that contain d- or f-elements do retain unquenched spins. Salts of such elements often show paramagnetic behavior but at low enough temperatures the magnetic moments may order. It is not uncommon to call such materials 'paramagnets', when referring to their paramagnetic behavior above their Curie or Nel-points, particularly if such temperatures are very low or have never been properly measured. Even for iron it is not uncommon to say that iron becomes a paramagnet above its relatively high Curie-point. In that case the Curie-point is Idealized CurieWeiss behavior; N.B. TC=, but TN is not . Paramagnetic seen as a phase transition between a regimes are denoted by solid lines. Close to TN or TC the behavior usually deviates ferromagnet and a 'paramagnet'. The word from ideal. paramagnet now merely refers to the linear response of the system to an applied field, the temperature dependence of which requires an amended version of Curie's law, known as the CurieWeiss law:

Paramagnetism

124

This amended law includes a term that describes the exchange interaction that is present albeit overcome by thermal motion. The sign of depends on whether ferro- or antiferromagnetic interactions dominate and it is seldom exactly zero, except in the dilute, isolated cases mentioned above. Obviously, the paramagnetic CurieWeiss description above TN or TC is a rather different interpretation of the word 'paramagnet' as it does not imply the absence of interactions, but rather that the magnetic structure is random in the absence of an external field at these sufficiently high temperatures. Even if is close to zero this does not mean that there are no interactions, just that the aligning ferro- and the anti-aligning antiferromagnetic ones cancel. An additional complication is that the interactions are often different in different directions of the crystalline lattice (anisotropy), leading to complicated magnetic structures once ordered. Randomness of the structure also applies to the many metals that show a net paramagnetic response over a broad temperature range. They do not follow a Curie type law as function of temperature however, often they are more or less temperature independent. This type of behavior is of an itinerant nature and better called Pauli-paramagnetism, but it is not unusual to see e.g. the metal aluminium called a 'paramagnet', even though interactions are strong enough to give this element very good electrical conductivity.

Superparamagnets
Some materials show induced magnetic behavior that follows a Curie type law but with exceptionally large values for the Curie constants. These materials are known as superparamagnets. They are characterized by a strong ferromagnetic or ferrimagnetic type of coupling into domains of a limited size that behave independently from one another. The bulk properties of such a system resembles that of a paramagnet, but on a microscopic level they are ordered. The materials do show an ordering temperature above which the behavior reverts to ordinary paramagnetism (with interaction). Ferrofluids are a good example, but the phenomenon can also occur inside solids, e.g., when dilute paramagnetic centers are introduced in a strong itinerant medium of ferromagnetic coupling such as when Fe is substituted in TlCu2Se2 or the alloy AuFe. Such systems contain ferromagnetically coupled clusters that freeze out at lower temperatures. They are also called mictomagnets.

References
[1] G. L. Miessler and D. A. Tarr Inorganic Chemistry 3rd Ed, Pearson/Prentice Hall publisher, ISBN 0-13-035471-6. [2] Nave, Carl L. "Magnetic Properties of Solids" (http:/ / hyperphysics. phy-astr. gsu. edu/ Hbase/ tables/ magprop. html). HyperPhysics. . Retrieved 2008-11-09. [3] J. Jensen and A. R. MacKintosh, "Rare Earth Magnetism" (http:/ / www2. nbi. ku. dk/ page40667. htm). . Retrieved 2009-07-12., (Clarendon Press, Oxford: 1991). [4] A. F. Orchard, Magnetochemistry, (Oxford University Press: 2003).

General reference texts


Charles Kittel, Introduction to Solid State Physics (Wiley: New York, 1996). Neil W. Ashcroft and N. David Mermin, Solid State Physics (Harcourt: Orlando, 1976). John David Jackson, Classical Electrodynamics (Wiley: New York, 1999).

Paramagnetism

125

External links
http://www.ndt-ed.org/EducationResources/CommunityCollege/MagParticle/Physics/MagneticMatls.htm

Electron magnetic dipole moment


In atomic physics, the electron magnetic dipole moment is the magnetic moment of an electron caused by its intrinsic property of spin.

Magnetic moment of an electron


The electron is a charged particle of charge (e), where e is the elementary charge. Its angular momentum comes from two types of rotation: spin and orbital motion. From classical electrodynamics, a rotating electrically charged body creates a magnetic dipole with magnetic poles of equal magnitude but opposite polarity. This analogy holds as an electron indeed behaves like a tiny bar magnet. One consequence is that an external magnetic field exerts a torque on the electron magnetic moment depending on its orientation with respect to the field. If the electron is visualized as a classical charged particle literally rotating about an axis with angular momentum L, its magnetic dipole moment is given by:

where me is the electron rest mass. Note that the angular momentum L in this equation may be the spin angular momentum, the orbital angular momentum, or the total angular momentum. It turns out the classical result is off by a proportional factor for the spin magnetic moment. As a result, the classical result is corrected by multiplying it with a dimensionless correction factor g is known as the g-factor;

It is usual to express the magnetic moment in terms of the reduced Planck constant and the Bohr magneton B:

since the magnetic moment is quantized in units of B, correspondingly the angular momentum is quantized in units of .

Spin magnetic dipole moment


The spin magnetic moment is intrinsic for an electron.[1] It is:

Here S is the electron spin angular momentum. The spin g-factor is approximately two: gs 2. The magnetic moment of an electron is approximately twice what it should be in classical mechanics. The factor of two implies that the electron appears to be twice as effective in producing a magnetic moment as the corresponding classical charged body. The spin magnetic dipole moment is approximately one B because g 2 and the electron is a spin one-half particle: S = /2.

The z component of the electron magnetic moment is:

Electron magnetic dipole moment where mS is the spin quantum number. Note that is a negative constant multiplied by the spin, so the magnetic moment is antiparallel to the spin angular momentum. The spin g-factor gs = 2 comes from the Dirac equation, a fundamental equation connecting the electron's spin with its electromagnetic properties. Reduction of the Dirac equation for an electron in a magnetic field to its non-relativistic limit yields the Schrdinger equation with a correction term which takes account of the interaction of the electron's intrinsic magnetic moment with the magnetic field giving the correct energy. For the electron spin, the most accurate value for the spin g-factor has been experimentally determined to have the value 2.00231930419922 (1.5 1012).[2] Note that it is only two thousandths larger than the value from Dirac equation. The small correction is known as the anomalous magnetic dipole moment of the electron; it arises from the electron's interaction with virtual photons in quantum electrodynamics. In fact, one famous triumph of the Quantum Electrodynamics theory is the accurate prediction of the electron g-factor. The most accurate value for the electron magnetic moment is -928.476377 1026 0.000023 1026 JT1.[3]

126

Orbital magnetic dipole moment


The revolution of an electron around an axis through another object, such as the nucleus, gives rise to the orbital magnetic dipole moment. Suppose that the angular momentum for the orbital motion is L. Then the orbital magnetic dipole moment is:

Here gL is the electron orbital g-factor and B is the Bohr magneton. The value of gL is exactly equal to one, by a quantum-mechanical argument analogous to the derivation of the classical gyromagnetic ratio.

Total magnetic dipole moment


The total magnetic dipole moment resulting from both spin and orbital angular momenta of an electron is related to the total angular momentum J by a similar equation:

The g-factor gJ is known as the Land g-factor, which can be related to gL and gS by quantum mechanics. See Land g-factor for detials.

Example: hydrogen atom


For a hydrogen atom, an electron occupying the atomic orbital n, , m, the magnetic dipole moment is given by:

Here L is the orbital angular momentum, n, and m are the principal, azimuthal and magnetic quantum numbers respectively. The z-component of the orbital magnetic dipole moment for an electron with a magnetic quantum number m is given by:

Electron magnetic dipole moment

127

Electron spin in the Pauli and Dirac theories


The necessity of introducing half-integral spin goes back experimentally to the results of the SternGerlach experiment. A beam of atoms is run through a strong non-uniform magnetic field, which then splits into N parts depending on the intrinsic angular momentum of the atoms. It was found that for silver atoms, the beam was split in twothe ground state therefore could not be integral, because even if the intrinsic angular momentum of the atoms were as small as possible, 1, the beam would be split into 3 parts, corresponding to atoms with Lz = 1, 0, and +1. The conclusion is that silver atoms have net intrinsic angular momentum of 12. Pauli set up a theory which explained this splitting by introducing a two-component wave function and a corresponding correction term in the Hamiltonian, representing a semi-classical coupling of this wave function to an applied magnetic field, as so:

Here A is the magnetic potential and the electric potential representing the electromagnetic field, and = (x, y, z) are the Pauli matrices. On squaring out the first term, a residual interaction with the magnetic field is found, along with the usual classical Hamiltonian of a charged particle interacting with an applied field:

This Hamiltonian is now a 2 2 matrix, so the Schrdinger equation based on it must use a two-component wave function. Pauli had introduced the 2 2 sigma matrices as pure phenomenology Dirac now had a theoretical argument that implied that spin was somehow the consequence of incorporating relativity into quantum mechanics. On introducing the external electromagnetic 4-potential into the Dirac equation in a similar way, known as minimal coupling, it takes the form (in natural units = c = 1)

where are the gamma matrices (aka Dirac matrices) and i is the imaginary unit. A second application of the Dirac operator will now reproduce the Pauli term exactly as before, because the spatial Dirac matrices multiplied by i, have the same squaring and commutation properties as the Pauli matrices. What is more, the value of the gyromagnetic ratio of the electron, standing in front of Pauli's new term, is explained from first principles. This was a major achievement of the Dirac equation and gave physicists great faith in its overall correctness. The Pauli theory may be seen as the low energy limit of the Dirac theory in the following manner. First the equation is written in the form of coupled equations for 2-spinors with the units restored:

so

Assuming the field is weak and the motion of the electron non-relativistic, we have the total energy of the electron approximately equal to its rest energy, and the momentum reducing to the classical value,

and so the second equation may be written

which is of order v/c - thus at typical energies and velocities, the bottom components of the Dirac spinor in the standard representation are much suppressed in comparison to the top components. Substituting this expression into

Electron magnetic dipole moment the first equation gives after some rearrangement

128

The operator on the left represents the particle energy reduced by its rest energy, which is just the classical energy, so we recover Pauli's theory if we identify his 2-spinor with the top components of the Dirac spinor in the non-relativistic approximation. A further approximation gives the Schrdinger equation as the limit of the Pauli theory. Thus the Schrdinger equation may be seen as the far non-relativistic approximation of the Dirac equation when one may neglect spin and work only at low energies and velocities. This also was a great triumph for the new equation, as it traced the mysterious i that appears in it, and the necessity of a complex wave function, back to the geometry of space-time through the Dirac algebra. It also highlights why the Schrdinger equation, although superficially in the form of a diffusion equation, actually represents the propagation of waves. It should be strongly emphasized that this separation of the Dirac spinor into large and small components depends explicitly on a low-energy approximation. The entire Dirac spinor represents an irreducible whole, and the components we have just neglected to arrive at the Pauli theory will bring in new phenomena in the relativistic regime - antimatter and the idea of creation and annihilation of particles. In a general case (if a certain linear function of electromagnetic field does not vanish identically), three out of four components of the spinor function in the Dirac equation can be algebraically eliminated, yielding an equivalent fourth-order partial differential equation for just one component. Furthermore, this remaining component can be made real by a gauge transform.[4]

Notes
[1] A. Mahajan and A. Rangwala. Electricity and Magnetism (http:/ / books. google. com/ books?id=_tXrjggX7WwC& pg=PA419& lpg=PA419& dq="intrinsic+ dipole+ moment"+ and+ electron+ "Bohr+ magneton"& source=web& ots=87QUlLPdmD& sig=cmYr28QQJM75lI_ih4sS9UjGRE0), p. 419 (1989). Via Google Books. [2] http:/ / physics. nist. gov/ cgi-bin/ cuu/ Value?eqae|search_for=electron+ magnetic+ moment [3] http:/ / physics. nist. gov/ cgi-bin/ cuu/ Value?muem|search_for=magnetic+ moment+ electron [4] Source: Journal of Mathematical Physics, 52, 082303 (2011) (http:/ / jmp. aip. org/ resource/ 1/ jmapaq/ v52/ i8/ p082303_s1 or http:/ / akhmeteli. org/ wp-content/ uploads/ 2011/ 08/ JMAPAQ528082303_1. pdf )

Pascal's constants

129

Pascal's constants
Pascals constants are numbers used in the evaluation of the magnetic susceptibilities of coordination compounds. The magnetic susceptibility of a compound is the sum of the paramagnetic susceptibility associated with the unpaired electrons and the opposing diamagnetic susceptibility associated with electron pairs.[1] Typically, the paramagnetic susceptibility greatly exceeds in magnitude the diamagnetic susceptibility. Thus, the diamagnetic correction is not considered for many purposes. For more precise analyses, however, the diamagnetic corrections are calculated by summing the contributions from the components of the molecule. These group contributions are Pascals constants. This analysis assumes that these group contributions are identical in all molecules. In general, the magnitude of Pascals constants correlates with the number of electrons in the groups. Groups with extended pi-delocalization have larger diamagnetic corrections compared to related saturated ligands. These correction factors were first described by Pascal in 1910.[2] The values and the method of analysis have been revised several times. [1]

Representative Pascal Constants[3]


cation value anion value Li+ Na+ K+ Rb+ Cs+ NH4+ Mg2+ Fe2+ Co2+ Ni2+ Cu2+ Zn2+ Cd2+ Hg2+ Pb2+ -1.0 -6.8 F Cl -9.1 -23.4 -34.6 -50.6 -12.0 -13.0 -31.0

-14.9 Br -22.5 I -35 -40 -5 OHCNNCS-

-12.8 CO 2 -29.5 3 -12.8 NO 3 -12.8 SO 24 -12.8 ClO 3 -15 -24 ClO4-18.9 -40.1 -30.2 -32

BrO3- -38.8 -51.4 -51.9

-40.0 IO 3 -32.0 IO 4

Pascal's constants

130

Representative Pascal Constants[3]


ligands H2O NH3 pyridine ethylenediamine bipyridyl phenanthroline acetylacetonate glycinate value -13 -18 -49 -46.5 -105 -128 -52 -37

Triphenylphosphine -167

Representative Pascal Constants[3]


In covalent compounds value H B C C (ring) N (open chain) N (ring) O F Cl Br I -2.93 -7.0 -6.00 -6.24 -5.57 -4.61 -4.6 -6.3 -20.1 -30.6 -44.6

With organic molecules corrections are made for bond types other than C-C and C-H single bonds.

Correction Constants[3]
Bond type value C=C CC C=O COOH CN +5.5 +0.8 +6.3 -5.0 0.0

Pascal's constants

131

Notes
[1] Bain, Gordon A.; Berry , John F. (2008). "Diamagnetic Corrections and Pascals Constants". J. Chem. Educ. 85: 532. doi:10.1021/ed085p532. pdf (http:/ / www. jce. divched. org/ JCEDLib/ ChemInfo/ Inorganic/ pascal. html) [2] Pascal, P. Ann. Chim. Phys. 1910, volume 19, page 5 [3] values in CGS units, nn10-6/g atom

References

CurieWeiss law
The Curie-Weiss law describes the magnetic susceptibility of a ferromagnet in the paramagnetic region above the Curie point:

where C is a material-specific Curie constant, T is absolute temperature, measured in kelvins, and Tc is the Curie temperature, measured in kelvins. The law predicts a singularity in the susceptibility at T = Tc. Below this temperature the ferromagnet has a spontaneous magnetization. In many materials the Curie-Weiss law fails to describe the susceptibility in the immediate vicinity of the Curie point, since it is based on a mean-field approximation. Instead, there is a critical behavior of the form

with the critical exponent . However, at temperatures T >> Tc the expression of the Curie-Weiss law still holds, but with Tc replaced by a temperature that is somewhat higher than the actual Curie temperature. Some authors call the Weiss constant to distinguish it from the temperature of the actual Curie point.

References
Introduction to Solid State Physics 7th ed. (1996) by Charles Kittel

Curie's law

132

Curie's law
In a paramagnetic material the magnetization of the material is (approximately) directly proportional to an applied magnetic field. However, if the material is heated, this proportionality is reduced: for a fixed value of the field, the magnetization is (approximately) inversely proportional to temperature. This fact is encapsulated by Curie's law:

where is the resulting magnetisation is the magnetic field, measured in teslas is absolute temperature, measured in kelvins is a material-specific Curie constant. This relation was discovered experimentally (by fitting the results to a correctly guessed model) by Pierre Curie. It only holds for high temperatures, or weak magnetic fields. As the derivations below show, the magnetization saturates in the opposite limit of low temperatures, or strong fields.

Derivation with quantum statistical mechanics


A simple model of a paramagnet concentrates on the particles which compose it which do not interact with each other. Each particle has a magnetic moment given by . The energy of a magnetic moment in a magnetic field is given by

Two-state (spin-1/2) particles


To simplify the calculation, we are going to work with a 2-state particle: it may either align its magnetic moment with the magnetic field, or against it. So the only possible values of magnetic moment are then and . If so, then such a particle has only two possible energies

and

Magnetization of a paramagnet as a function of inverse temperature.

When one seeks the magnetization of a paramagnet, one is interested in the likelihood of a particle to align itself with the field. In other words, one seeks the expectation value of the magnetization :

where the probability of a configuration is given by its Boltzmann factor, and the partition function

provides the

necessary normalization for probabilities (so that the sum of all of them is unity.) The partition function of one particle is:

Curie's law

133

Therefore, in this simple case we have:

This is magnetization of one particle, the total magnetization of the solid is given by

The formula above is known as the Langevin paramagnetic equation. Pierre Curie found an approximation to this law which applies to the relatively high temperatures and low magnetic fields used in his experiments. Let's see what happens to the magnetization as we specialize it to large and small . As temperature increases and magnetic field decreases, the argument of hyperbolic tangent decreases. Another way to say this is

this is sometimes called the Curie regime. We also know that if so

, then

with a Curie constant given by fields, tends to a maximum value of with the field.

. Also, in the opposite regime of low temperatures or high , corresponding to all the particles being completely aligned

General case
When the particles have an arbitrary spin (any number of spin states), the formula is a bit more complicated. For this more general formula and its derivation, see the article: Brillouin function. As the spin approaches infinity, the formula for the magnetization approaches the classical value derived in the following section. At low magnetic fields or high temperature, however, a simple Curie law is again obtained (where angular momentum quantum number): is the total

Derivation with classical statistical mechanics


An alternative treatment applies when the paramagnetons are imagined to be classical, freely-rotating magnetic moments. In this case, their position will be determined by their angles in spherical coordinates, and the energy for one of them will be:

where

is the angle between the magnetic moment and the magnetic field (which we take to be pointing in the

coordinate.) The corresponding partition function is

We see there is no dependence on the

angle, and also we can change variables to

to obtain

Curie's law Now, the expected value of the component of the magnetization (the other two are seen to be null (due to integration over ), as they should) will be given by

134

To simplify the calculation, we see this can be written as a differentiation of

(This approach can also be used for the model above, but the calculation was so simple this is not so helpful.) Carrying out the derivation we find

where

is the Langevin function:

This function would appear to be singular for small In fact, its behavior for small arguments is opposite limit is likewise recovered.

, but it is not, since the two singular terms cancel each other. , so the Curie limit also applies, but with a Curie constant for large values of its argument, and the

three times smaller in this case. Similarly, the function saturates at

Applications
It is the basis of operation of magnetic thermometers, which are used to measure very low temperatures.

Curie constant
The Curie constant is a material-dependent property that relates a material's magnetic susceptibility to its temperature. The Curie constant, when expressed in SI units, is given by
[1]

where

is the permeability constant,

is the number of magnetic atoms (or molecules) per unit volume,

is the

Land g-factor, number and

(9.27400915e-24 J/T or Am2) is the Bohr magneton,

is the angular momentum quantum , the formula reduces to

is Boltzmann's constant. For a two-level system with magnetic moment

The constant is used in Curie's Law, which states that for a fixed value of a magnetic field, the magnetization of a material is (approximately) inversely proportional to temperature.

This equation was first derived by Pierre Curie. Because of the relationship between magnetic susceptibility , magnetization and applied magnetic field :

this shows that for a paramagnetic system of non-interacting magnetic moments, magnetization related to temperature (see Curie's Law).

is inversely

Curie constant

135

References
[1] Kittel, Charles. Introduction to Solid State Physics, 8th Edition. Wiley. pp.304. ISBN047141526X.

Boltzmann distribution
In chemistry, physics, and mathematics, the Boltzmann distribution (also called the Gibbs Distribution[1]) is a certain distribution function or probability measure for the distribution of the states of a system. The distribution was discovered in the context of classical statistical mechanics by J.W. Gibbs in 1901. It underpins the concept of the canonical ensemble, providing the underyling distribution. A special case of the Boltzmann distribution, used for describing the velocities of particles of a gas, is the MaxwellBoltzmann distribution. In more general mathematical settings, the Boltzmann distribution is also known as the Gibbs measure.

Definition
The Boltzmann distribution for the fractional number of particles Ni/N occupying a set of states i possessing energy Ei is:

where

is the Boltzmann constant, T is temperature (assumed to be a well-defined quantity),

is the degeneracy

(meaning, the number of levels having energy

; sometimes, the more general 'states' are used instead of levels, to

avoid using degeneracy in the equation), N is the total number of particles and Z(T) is the partition function.

Alternatively, for a single system at a well-defined temperature, it gives the probability that the system is in the specified state. The Boltzmann distribution applies only to particles at a high enough temperature and low enough density that quantum effects can be ignored, and the particles are obeying MaxwellBoltzmann statistics. (See that article for a derivation of the Boltzmann distribution.) The Boltzmann distribution is often expressed in terms of = 1/kT where is referred to as thermodynamic beta. The term or , which gives the (unnormalised) relative probability of a state, is called the Boltzmann factor and appears often in the study of physics and chemistry. When the energy is simply the kinetic energy of the particle

then the distribution correctly gives the MaxwellBoltzmann distribution of gas molecule speeds, previously predicted by Maxwell in 1859. The Boltzmann distribution is, however, much more general. For example, it also predicts the variation of the particle density in a gravitational field with height, if . In fact the distribution applies whenever quantum considerations can be ignored. In some cases, a continuum approximation can be used. If there are g(E)dE states with energy E to E+dE, then the Boltzmann distribution predicts a probability distribution for the energy:

Then g(E) is called the density of states if the energy spectrum is continuous. Classical particles with this energy distribution are said to obey MaxwellBoltzmann statistics.

Boltzmann distribution In the classical limit, i.e. at large values of or at small density of states when wave functions of

136

particles practically do not overlap both the BoseEinstein or FermiDirac distribution become the Boltzmann distribution.

Derivation
See MaxwellBoltzmann statistics.

References
[1] Landau, Lev Davidovich; and Lifshitz, Evgeny Mikhailovich (1980) [1976]. Statistical Physics. 5 (3 ed.). Oxford: Pergamon Press. ISBN0-7506-3372-7. Translated by J.B. Sykes and M.J. Kearsley. See section 28

External links
Derivation of the distribution for microstates of a system (http://theory.ph.man.ac.uk/~judith/stat_therm/ node67.html)

Brillouin and Langevin functions


Brillouin Function
The Brillouin function[1][2] is a special function defined by the following equation:

The function is usually applied (see below) in the context where x is a real variable and J is a positive integer or half-integer. In this case, the function varies from -1 to 1, approaching +1 as and -1 as . The function is best known for arising in the calculation of the magnetization of an ideal paramagnet. In particular, it describes the dependency of the magnetization on the applied magnetic field and the total angular momentum quantum number J of the microscopic magnetic moments of the material. The magnetization is given by:[1]

where is the number of atoms per unit volume, the g-factor, the Bohr magneton, is the ratio of the Zeeman energy of the magnetic moment in the external field to the thermal energy :

is the Boltzmann constant and

the temperature. given in Tesla stands for magnetic induction, is the permeability of vacuum. , where is the

Note that in the SI system of units

applied magnetic field given in A/m and


Click "show" to see a derivation of this law:

Brillouin and Langevin functions


[1] A derivation of this law describing the magnetization of an ideal paramagnet is as follows. Let z be the direction of the magnetic field. The z-component of the angular momentum of each magnetic moment (a.k.a. the azimuthal quantum number) can take on one of the 2J+1 possible values -J,-J+1,...,+J. Each of these has a different energy, due to the external field B: The energy associated with quantum number m is

137

(where g is the g-factor, B is the Bohr magneton, and x is as defined in the text above). The relative probability of each of these is given by the Boltzmann factor:

where Z (the partition function) is a normalization constant such that the probabilities sum to unity. Calculating Z, the result is: . All told, the expectation value of the azimuthal quantum number m is . The denominator is a geometric series and the numerator is a type of arithmetic-geometric series some algebra, the result turns out to be [3] , so the series can be explicitly summed. After

With N magnetic moments per unit volume, the magnetization density is .

Langevin Function
In the classical limit, the moments can be continuously aligned in the field and can assume all values ( ). The Brillouin function is then simplified into the Langevin function, named after Paul Langevin:

For small values of x, the Langevin function can be approximated by a truncation of its Taylor series:

An alternative better behaved approximation can be derived from the Lambert's continued fraction expansion of tanh(x):

Langevin function (red line), compared with

(blue line).

For small enough x, both approximations are numerically better than a direct evaluation of the actual analytical expression, since the later suffers from Loss of significance. The inverse Langevin function can be approximated to within 5% accuracy by the formula[4]

valid on the whole interval (-1, 1). For small values of x, better approximations are the Pad approximant

Brillouin and Langevin functions

138

and the Taylor series[5]

High Temperature Limit


When Curie's law: i.e. when is small, the expression of the magnetization can be approximated by the

where magnetons.

is a constant. One can note that

is the effective number of Bohr

High Field Limit


When , the Brillouin function goes to 1. The magnetization saturates with the magnetic moments completely aligned with the applied field:

References
[1] C. Kittel, Introduction to Solid State Physics (8th ed.), pages 303-4 ISBN 978-0471415268 [2] Darby, M.I. (1967). "Tables of the Brillouin function and of the related function for the spontaneous magnetization". Brit. J. Appl. Phys. 18 (10): 14151417. Bibcode1967BJAP...18.1415D. doi:10.1088/0508-3443/18/10/307 [3] http:/ / planetmath. org/ encyclopedia/ ArithmeticGeometricSeries. html [4] Cohen, A. (1991). "A Pad approximant to the inverse Langevin function". Rheologica Acta 30 (3): 270273. doi:10.1007/BF00366640. [5] Johal, A. S.; Dunstan, D. J. (2007). "Energy functions for rubber from microscopic potentials". Journal of Applied Physics 101 (8): 084917. Bibcode2007JAP...101h4917J. doi:10.1063/1.2723870.

139

Paramagnetic Resonance
Gyromagnetic ratio
In physics, the gyromagnetic ratio (also sometimes known as the magnetogyric ratio in other disciplines) of a particle or system is the ratio of its magnetic dipole moment to its angular momentum, and it is often denoted by the symbol , gamma. Its SI units are radian per second per tesla (rad s1T -1) or, equivalently, coulomb per kilogram (Ckg1). The term "gyromagnetic ratio" is sometimes used[1] as a synonym for a different but closely related quantity, the g-factor. The g-factor, unlike the gyromagnetic ratio, is dimensionless. For more on the g-factor, see below, or see the article g-factor.

Gyromagnetic ratio and Larmor precession


Any free system with a constant gyromagnetic ratio, such as a rigid system of charges, a nucleus, or an electron, when placed in an external magnetic field B (measured in teslas) that is not aligned with its magnetic moment, will precess at a frequency f (measured in hertz), that is proportional to the external field: . For this reason, values of /(2), in units of hertz per tesla (Hz/T), are often quoted instead of . This relationship also explains an apparent contradiction between the two equivalent terms, gyromagnetic ratio versus magnetogyric ratio: whereas it is a ratio of a magnetic property (i.e. dipole moment) to a gyric (rotational, from Greek: , "turn") property (i.e. angular momentum), it is also, at the same time, a ratio between the angular precession frequency (another gyric property) = 2f and the magnetic field.

Gyromagnetic ratio for a classical rotating body


Consider a charged body rotating about an axis of symmetry. According to the laws of classical physics, it has both a magnetic dipole moment and an angular momentum due to its rotation. It can be shown that as long as its charge and mass are distributed identically (e.g., both distributed uniformly), its gyromagnetic ratio is

where q is its charge and m is its mass. The derivation of this relation is as follows: It suffices to demonstrate this for an infinitesimally narrow circular ring within the body, as the general result follows from an integration. Suppose the ring has radius r, area A = r2, mass m, charge q, and angular momentum L=mvr. Then the magnitude of the magnetic dipole moment is

as desired.

Gyromagnetic ratio

140

Gyromagnetic ratio for an isolated electron


An isolated electron has an angular momentum and a magnetic moment resulting from its spin. While an electron's spin is sometimes visualized as a literal rotation about an axis, it is in fact a fundamentally different, quantum-mechanical phenomenon[2] with no true analogue in classical physics. Consequently, there is no reason to expect the above classical relation to hold. In fact it does not, giving the wrong result by a dimensionless factor called the electron g-factor, denoted ge (or just g when there is no risk of confusion):

where B is the Bohr magneton. As mentioned above, in classical physics one would expect the g-factor to be . However in the framework of relativistic quantum mechanics,

where

is the fine-structure constant. Here the small corrections to the relativistic result

come

from the quantum field theory. Experimentally, the electron g-factor has been measured to twelve decimal places:[3] The electron gyromagnetic ratio is given by NIST[4] as

The g-factor and are in excellent agreement with theory; see Precision tests of QED for details.

Gyromagnetic factor as a consequence of relativity


Since a gyromagnetic factor equal to 2 follows from the Dirac's equation it is a frequent misconception to think that a g-factor 2 is a consequence of relativity; it is not. The factor 2 can be obtained from the linearization of both the Schrdinger equation and the relativistic Klein-Gordon equation (which leads to Dirac's). In both cases a 4-spinor is obtained and for both linearizations the g-factor is found to be equal to 2; Therefore, the factor 2 is a consequence of the wave equation dependency on the first (and not the second) derivatives with respect to space and time.[5]

Gyromagnetic ratio for a nucleus


Protons, neutrons, and many nuclei carry nuclear spin, which gives rise to a gyromagnetic ratio as above. The ratio is conventionally written in terms of the proton mass and charge, even for neutrons and for other nuclei, for the sake of simplicity and consistency. The formula is:

where

is the nuclear magneton, and g is the g-factor of the nucleon or nucleus in question.

The gyromagnetic ratio of a nucleus is particularly important because of the role it plays in Nuclear Magnetic Resonance (NMR) and Magnetic Resonance Imaging (MRI). These procedures rely on the fact that nuclear spins precess in a magnetic field at a rate called the Larmor frequency, which is simply the product of the gyromagnetic ratio with the magnetic field strength. Approximate values for some common nuclei are given in the Table below.[6] [7]

Gyromagnetic ratio

141

Nucleus / 106 rad s1 T 1 /2 / MHz T 1


1

H H He Li C N N O F Na P Xe

267.513 41.065 -203.789 103.962 67.262 19.331 -27.116 -36.264 251.662 70.761 108.291 -73.997

42.576 6.536 -32.434 16.546 10.705 3.077 -4.316 -5.772 40.053 11.262 17.235 -11.777

13

14

15

17

19

23

31

129

References
Note 1 note
Note 1: Marc Knecht, The Anomalous Magnetic Moments of the Electron and the Muon [8], Poincar Seminar (Paris, Oct. 12, 2002), published in : Duplantier, Bertrand; Rivasseau, Vincent (Eds.) ; Poincar Seminar 2002, Progress in Mathematical Physics 30, Birkhuser (2003), ISBN 3-7643-0579-7.

General note
[1] For example, see: D.C. Giancoli, Physics for Scientists and Engineers, 3rd ed., page 1017. Or see: P.A. Tipler and R.A. Llewellyn, Modern Physics, 4th ed., page 309. [2] S J Brodsky, V A Franke, J R Hiller, G McCartor, S A Paston and E V Prokhvatilov (2004). "A nonperturbative calculation of the electron's magnetic moment". Nuclear Physics B 703 (12): 333362. arXiv:hep-ph/0406325. Bibcode2004NuPhB.703..333B. doi:10.1016/j.nuclphysb.2004.10.027. [3] B Odom, D Hanneke, B D'Urso and G Gabrielse (2006). "New measurement of the electron magnetic moment using a one-electron quantum cyclotron". Physical Review Letters 97 (3): 030801. Bibcode2006PhRvL..97c0801O. doi:10.1103/PhysRevLett.97.030801. PMID16907490. [4] NIST (http:/ / physics. nist. gov/ cgi-bin/ cuu/ Value?gammae). Note that NIST puts a positive sign on the quantity; however, to be consistent with the formulas in this article, a negative sign is put on here. Indeed, many references say that <0 for an electron; for example, Weil and Bolton, Electron Paramagnetic Resonance (Wiley 2007), page 578 [5] Walter Greiner Quantum Mechanics: An Introduction, Springer Verlag [6] M A Bernstein, K F King and X J Zhou (2004). Handbook of MRI Pulse Sequences. San Diego: Elsevier Academic Press. p.960. ISBN0-1209-2861-2. [7] R C Weast, M J Astle, ed. (1982). Handbook of Chemistry and Physics. Boca Raton: CRC Press. p.E66. ISBN0-8493-0463-6. [8] http:/ / parthe. lpthe. jussieu. fr/ poincare/ textes/ octobre2002/ Knecht. ps

Electron paramagnetic resonance

142

Electron paramagnetic resonance


Electron paramagnetic resonance (EPR) or electron spin resonance (ESR) spectroscopy is a technique for studying chemical species that have one or more unpaired electrons, such as organic and inorganic free radicals or inorganic complexes possessing a transition metal ion. The basic physical concepts of EPR are analogous to those of nuclear magnetic resonance (NMR), but it is electron spins that are excited instead of spins of atomic nuclei. Because most stable molecules have all their electrons paired, the EPR technique is less widely used than NMR. However, this limitation to paramagnetic species also means that the EPR technique is one of great specificity, since ordinary chemical solvents and matrices do not give rise to EPR spectra. EPR was first observed in Kazan State University by Soviet physicist Yevgeny Zavoisky in 1944, and was developed independently at the same time by Brebis Bleaney at the University of Oxford.

Theory
Origin of an EPR signal
Every electron has a magnetic moment and spin quantum number , with magnetic components and

. In the presence of an external magnetic field with strength , the electron's magnetic moment aligns itself either parallel ( ) or antiparallel ( ) to the field, each alignment

having a specific energy (see the Zeeman EPR spectrometer effect). The parallel alignment corresponds to the lower energy state, and the separation between it and the upper state is , where is the electron's so-called g-factor (see also the Land g-factor) and strength, as shown in the diagram below.

is the Bohr

magneton. This equation implies that the splitting of the energy levels is directly proportional to the magnetic field's

Electron paramagnetic resonance An unpaired electron can move between the two energy levels by either absorbing or emitting electromagnetic radiation of energy such that the resonance condition, , is obeyed. Substituting in and leads to the fundamental equation of EPR spectroscopy: . Experimentally, this equation permits a large combination of frequency and magnetic field values, but the great majority of EPR measurements are made with microwaves in the 900010000MHz (910GHz) region, with fields corresponding to about 3500 G (0.35 T). See below for other field-frequency combinations. In principle, EPR spectra can be generated by either varying the photon frequency incident on a sample while holding the magnetic field constant or doing the reverse. In practice, it is usually the frequency that is kept fixed. A collection of paramagnetic centers, such as free radicals, is exposed to microwaves at a fixed frequency. By increasing an external magnetic field, the gap between the and energy states is widened until it matches the energy of the microwaves, as represented by the double-arrow in the diagram above. At this point the unpaired electrons can move between their two spin states. Since there typically are more electrons in the lower state, due to the Maxwell-Boltzmann distribution (see below), there is a net absorption of energy, and it is this absorption that is monitored and converted into a spectrum. As an example of how can be used, consider the case of a free electron, which has = 2.0023,[1] and the simulated spectrum shown at the right in two different forms. For the microwave frequency of 9388.2MHz, the predicted resonance position is a magnetic field of about = 0.3350 tesla = 3350 gauss, as shown. Note that while two forms of the same spectrum are presented in the figure, most EPR spectra are recorded and published only as first derivatives. Because of electron-nuclear mass differences, the magnetic moment of an electron is substantially larger than the corresponding quantity for any nucleus, so that a much higher electromagnetic frequency is needed to bring about a spin resonance with an electron than with a nucleus, at identical magnetic field strengths. For example, for the field of 3350 G shown at the right, spin resonance occurs near 9388.2MHz for an electron compared to only about 14.3MHz for 1H nuclei. (For NMR spectroscopy, the corresponding resonance equation is where and depend on the nucleus under study.)

143

Maxwell-Boltzmann distribution
In practice, EPR samples consist of collections of many paramagnetic species, and not single isolated paramagnetic centers. If the population of radicals is in thermodynamic equilibrium, its statistical distribution is described by the Maxwell-Boltzmann equation

where constant, and

is the number of paramagnetic centers occupying the upper energy state, is the temperature in kelvins. At 298 K, X-band microwave frequencies (

is the Boltzmann 9.75GHz) give

0.998, meaning that the upper energy level has a smaller population than the lower one. Therefore, transitions from the lower to the higher level are more probable than the reverse, which is why there is a net

Electron paramagnetic resonance absorption of energy. The sensitivity of the EPR method (i.e., the minimum number of detectable spins frequency according to ) depends on the photon

144

where With

is a constant, and

is the sample's volume, ~

is the unloaded quality factor of the microwave cavity is the microwave power in the spectrometer cavity. ~ , where 1.5. In practice, can , i.e.,

(sample chamber),

is the cavity filling coefficient, and

being constants,

change varying from 0.5 to 4.5 depending on spectrometer characteristics, resonance conditions, and sample size. In other words, the higher the spectrometer frequency the lower the detection limit ( sensitivity. ), meaning greater

Spectral parameters
In real systems, electrons are normally not solitary, but are associated with one or more atoms. There are several important consequences of this: 1. An unpaired electron can gain or lose angular momentum, which can change the value of its g-factor, causing it to differ from . This is especially significant for chemical systems with transition-metal ions. 2. If an atom with which an unpaired electron is associated has a non-zero nuclear spin, then its magnetic moment will affect the electron. This leads to the phenomenon of hyperfine coupling, analogous to J-coupling in NMR, splitting the EPR resonance signal into doublets, triplets and so forth. 3. Interactions of an unpaired electron with its environment influence the shape of an EPR spectral line. Line shapes can yield information about, for example, rates of chemical reactions. 4. The g-factor and hyperfine coupling in an atom or molecule may not be the same for all orientations of an unpaired electron in an external magnetic field. This anisotropy depends upon the electronic structure of the atom or molecule (e.g., free radical) in question, and so can provide information about the atomic or molecular orbital containing the unpaired electron.

The g factor
Knowledge of the g-factor can give information about a paramagnetic center's electronic structure. An unpaired electron responds not only to a spectrometer's applied magnetic field but also to any local magnetic fields of atoms or molecules. The effective field experienced by an electron is thus written

where

includes the effects of local fields (

can be positive or negative). Therefore, the

resonance condition (above) is rewritten as follows: The quantity is denoted and called simply the -factor, so that the final resonance equation becomes

This last equation is used to determine resonance occurs. If does not equal

in an EPR experiment by measuring the field and the frequency at which the implication is that the ratio of the unpaired electron's spin magnetic

moment to its angular momentum differs from the free electron value. Since an electron's spin magnetic moment is constant (approximately the Bohr magneton), then the electron must have gained or lost angular momentum through spin-orbit coupling. Because the mechanisms of spin-orbit coupling are well understood, the magnitude of the change gives information about the nature of the atomic or molecular orbital containing the unpaired electron.

Electron paramagnetic resonance In general, the g factor is not a number but a second-rank tensor represented by nine numbers arranged in a 33 matrix. The principal axes of this tensor are determined by the local fields, for example, by the local atomic arrangement around the unpaired spin in a solid or in a molecule. Choosing an appropriate coordinate system (say, x,y,z) allows to "diagonalize" this tensor thereby reducing the maximum number of its components from nine to three, gxx, gyy and gzz. For a single spin experiencing only Zeeman interaction with an external magnetic field, the position of the EPR resonance is given by the expression gxxBx + gyyBy + gzzBz. Here Bx, By and Bz are the components of the magnetic field vector in the coordinate system (x,y,z); their magnitudes change as the field is rotated, so as the frequency of the resonance. For a large ensemble of randomly oriented spins, the EPR spectrum consists three peaks of characteristic shape at frequencies gxxB0, gyyB0 and gzzB0: the low-frequency peak is positive in first-derivative spectra, the high-frequency peak is negative, and the central peak is bipolar. Such situation is commonly observed in powders and the spectra are therefore called "powder-pattern spectra". In crystals, the number of EPR lines is determined by the number of crystallographically equivalent orientations of the EPR spin (called "EPR center").

145

Hyperfine coupling
Since the source of an EPR spectrum is a change in an electron's spin state, it might be thought that all EPR spectra for a single electron spin would consist of one line. However, the interaction of an unpaired electron, by way of its magnetic moment, with nearby nuclear spins, results in additional allowed energy states and, in turn, multi-lined spectra. In such cases, the spacing between the EPR spectral lines indicates the degree of interaction between the unpaired electron and the perturbing nuclei. The hyperfine coupling constant of a nucleus is directly related to the spectral line spacing and, in the simplest cases, is essentially the spacing itself. Two common mechanisms by which electrons and nuclei interact are the Fermi contact interaction and by dipolar interaction. The former applies largely to the case of isotropic interactions (independent of sample orientation in a magnetic field) and the latter to the case of anisotropic interactions (spectra dependent on sample orientation in a magnetic field). Spin polarization is a third mechanism for interactions between an unpaired electron and a nuclear spin, being especially important for -electron organic radicals, such as the benzene radical anion. The symbols "a" or "A" are used for isotropic hyperfine coupling constants while "B" is usually employed for anisotropic hyperfine coupling constants.[2] In many cases, the isotropic hyperfine splitting pattern for a radical freely tumbling in a solution (isotropic system) can be predicted. For a radical having M equivalent nuclei, each with a spin of I, the number of EPR lines expected is 2MI+1. As an example, the methyl radical, CH3, has three 1H nuclei each with I=1/2, and so the number of lines expected is 2MI+ 1= 2(3)(1/2)+ 1= 4, which is as observed. For a radical having M1 equivalent nuclei, each with a spin of I1, and a group of M2 equivalent nuclei, each with a spin of I2, the number of lines expected is (2M1I1+ 1) (2M2I2+ 1). As an example, the methoxymethyl radical, H2C(OCH3), has two equivalent 1H nuclei each with I= 1/2 and three equivalent 1H nuclei each with I= 1/2, and so the number of lines expected is (2M1I1+ 1) (2M2I2+ 1) = [2(2)(1/2) + 1][2(3)(1/2) + 1] = [3][4] = 12, again as observed.

Electron paramagnetic resonance

146

The above can be extended to predict the number of lines for any number of nuclei. While it is easy to predict the number of lines a radical's EPR spectrum should show, the reverse problem, unraveling a complex multi-line EPR spectrum and assigning the various spacings to specific nuclei, is more difficult. In the oft-encountered case of I= 1/2 nuclei (e.g., 1H, 19F, 31P), the line intensities Simulated EPR spectrum of the CH3 radical produced by a population of radicals, each possessing M equivalent nuclei, will follow Pascal's triangle. For example, the spectrum at the right shows that the three 1H nuclei of the CH3 radical give rise to 2MI+ 1= 2(3)(1/2)+ 1= 4 lines with a 1:3:3:1 ratio. The line spacing gives a hyperfine coupling constant of aH = 23 G for each of the three 1H nuclei. Note again that the lines in this spectrum are first derivatives of absorptions. As a second example, consider the methoxymethyl radical, H2C(OCH3). The two equivalent methyl hydrogens will give an overall 1:2:1 EPR pattern, each component of which is further split by the three methoxy hydrogens into a 1:3:3:1 pattern to give a total of 34= 12 lines, a triplet of quartets. A simulation of the observed EPR spectrum is shown at the right, and agrees with the 12-line prediction Simulated EPR spectrum of the H2C(OCH3) radical and the expected line intensities. Note that the smaller coupling constant (smaller line spacing) is due to the three methoxy hydrogens, while the larger coupling constant (line spacing) is from the two hydrogens bonded directly to the carbon atom bearing the unpaired electron. It is often the case that coupling constants decrease in size with distance from a radical's unpaired electron, but there are some notable exceptions, such as the ethyl radical (CH2CH3).

Resonance linewidth definition


Resonance linewidths are defined in terms of the magnetic induction B, and its corresponding units, and are measured along the x axis of an EPR spectrum, from a line's center to a chosen reference point of the line. These defined widths are called halfwidths and possess some advantages: for asymmetric lines values of left and right halfwidth can be given. The halfwidth is the distance measured from the line's center to the point in which absorption value has half of maximal absorption value in the center of resonance line. First inclination width is a distance from center of the line to the point of maximal absorption curve inclination. In practice, a full definition of linewidth is used. For symmetric lines, halfwidth , and full inclination width

Electron paramagnetic resonance

147

Pulsed EPR
The dynamics of electron spins are best studied with pulsed measurements.[3] Microwave pulses typically 10100 ns long are used to control the spins in the Bloch sphere. The spin-lattice relaxation time can be measured with an inversion recovery experiment. As with pulsed NMR, the Hahn echo is central to many pulsed EPR experiments. A Hahn echo decay experiment can be used to measure the dephasing time, as shown in the animation below. The size of the echo is recorded for different spacings of the two pulses. This reveals the decoherence, which is not refocused by the pulse. In simple cases, an exponential decay is measured, which is described by the time.

Applications
EPR spectroscopy is used in various branches of science, such as chemistry and physics, for the detection and identification of free radicals and paramagnetic centers such as F centers. EPR is a sensitive, specific method for studying both radicals formed in chemical reactions and the reactions themselves. For example, when frozen water (solid H2O) is decomposed by exposure to high-energy radiation, radicals such as H, OH, and HO2 are produced. Such radicals can be identified and studied by EPR. Organic and inorganic radicals can be detected in electrochemical systems and in materials exposed to UV light. In many cases, the reactions to make the radicals and the subsequent reactions of the radicals are of interest, while in other cases EPR is used to provide information on a radical's geometry and the orbital of the unpaired electron. Medical and biological applications of EPR also exist. Although radicals are very reactive, and so do not normally occur in high concentrations in biology, special reagents have been developed to spin-label molecules of interest. These reagents are particularly useful in biological systems. Specially-designed nonreactive radical molecules can attach to specific sites in a biological cell, and EPR spectra can then give information on the environment of these so-called spin-label or spin-probes. A type of dosimetry system has been designed for reference standards and routine use in medicine, based on EPR signals of radicals from irradiated polycrystalline -alanine(the alanine deamination radical, the hydrogen abstraction radical, and the (CO-(OH))=C(CH3)NH2+ radical) . This method is suitable for measuring gamma and x-rays, electrons, protons, and high-linear energy transfer (LET) radiation of doses in the 1 Gy to 100 kGy range.[4] EPR spectroscopy can be applied only to systems in which the balance between radical decay and radical formation keeps the free-radicals concentration above the detection limit of the spectrometer used. This can be a particularly severe problem in studying reactions in liquids. An alternative approach is to slow down reactions by studying samples held at cryogenic temperatures, such as 77 K (liquid nitrogen) or 4.2 K (liquid helium). An example of this work is the study of radical reactions in single crystals of amino acids exposed to x-rays, work that sometimes leads to activation energies and rate constants for radical reactions. The study of radiation-induced free radicals in biological substances (for cancer research) poses the additional problem that tissue contains water, and water (due to its electric dipole moment) has a strong absorption band in the microwave region used in EPR spectrometers. EPR also has been used by archaeologists for the dating of teeth. Radiation damage over long periods of time creates free radicals in tooth enamel, which can then be examined by EPR and, after proper calibration, dated. Alternatively, material extracted from the teeth of people during dental procedures can be used to quantify their cumulative exposure to ionizing radiation. People exposed to radiation from the Chernobyl disaster have been examined by this method.[5][6] Radiation-sterilized foods have been examined with EPR spectroscopy, the aim being to develop methods to determine if a particular food sample has been irradiated and to what dose. Because of its high sensitivity, EPR was used recently to measure the quantity of energy used locally during a mechanochemical milling process.[7]

Electron paramagnetic resonance EPR spectroscopy has been used to measure properties of crude oil, in particular asphaltene and vanadium content. EPR measurement of asphaltene content is a function of spin density and solvent polarity. Prior work dating to the 1960s has demonstrated the ability to measure vanadium content to sub-ppm levels.

148

High-field high-frequency measurements


High-field high-frequency EPR measurements are sometimes needed to detect subtle spectroscopic details. However, for many years the use of electromagnets to produce the needed fields above 1.5 T was impossible, due principally to limitations of traditional magnet materials. The first multifunctional millimeter EPR spectrometer with a superconducting solenoid was described in the early 1970s by Prof. Y. S. Lebedev's group (Russian Institute of Chemical Physics, Moscow) in collaboration with L. G. Oranski's group (Ukrainian Physics and Technics Institute, Donetsk), which began working in the Institute of Problems of Chemical Physics, Chernogolovka around 1975.[8] Two decades later, a W-band EPR spectrometer was produced as a small commercial line by the German Bruker Company, initiating the expansion of W-band EPR techniques into medium-sized academic laboratories. Today there still are only a few scientific centers in the world capable of high-field high-frequency EPR; among them are the Grenoble High Magnetic Field Laboratory in Grenoble, France, the Physics Department in Freie Universitt Berlin, the National High Magnetic Field Laboratory in Tallahassee, US, the National Center for Advanced ESR Technology (ACERT) at Cornell University in Ithaca, US, the Department of Physiology, and Biophysics at Albert Einstein College of Medicine, Bronx, NY, the HLD and IFW in Dresden, Germany, the Institute of Physics of Complex Matter in Lausanne in Switzerland, and the Institute of Physics of the Leiden University, Netherlands.
Waveband L 300 1 S 100 3 C 75 4 X 30 10 P 20 15 K Q U 6 50 V E W F D 2.1 1.6 J 1.1 0.83 360

12.5 8.5 24 35

4.6 4 65 75

3.2 2.7 95

111 140 190 285 4.9 6.8

0.03 0.11 0.14 0.33 0.54 0.86 1.25 1.8 2.3 2.7 3.5 3.9

10.2 12.8

The EPR waveband is stipulated by the frequency or wavelength of a spectrometer's microwave source (see Table). EPR experiments often are conducted at X and, less commonly, Q bands, mainly due to the ready availability of the necessary microwave components (which originally were developed for radar applications). A second reason for widespread X and Q band measurements is that electromagnets can reliably generate fields up to about 1 tesla. However, the low spectral resolution over g-factor at these wavebands limits the study of paramagnetic centers with comparatively low anisotropic magnetic parameters. Measurements at > 40GHz, in the millimeter wavelength region, offer the following advantages: 1. EPR spectra are simplified due to the reduction of second-order effects at high fields. 2. Increase in orientation selectivity and sensitivity in the investigation of disordered systems. 3. The informativity and precision of pulse methods, e.g., ENDOR also increase at high magnetic fields.

EPR spectra of a nitroxide radical as a function of frequency. Note the improvement in [8] resolution from left to right.

4. Accessibility of spin systems with larger zero-field splitting due to the larger microwave quantum energy h

5. The higher spectral resolution over g-factor, which increases with irradiation frequency and external magnetic field B0. This is used to investigate the structure, polarity, and dynamics of radical microenvironments in spin-modified organic and biological systems through the spin label and probe method. The figure shows how spectral resolution improves with increasing frequency.

Electron paramagnetic resonance 6. Saturation of paramagnetic centers occurs at a comparatively low microwave polarizing field B1, due to the exponential dependence of the number of excited spins on the radiation frequency . This effect can be successfully used to study the relaxation and dynamics of paramagnetic centers as well as of superslow motion in the systems under study. 7. The cross-relaxation of paramagnetic centers decreases dramatically at high magnetic fields, making it easier to obtain more-precise and more-complete information about the system under study.[8]

149

References
[1] Odom, B.; Hanneke, D.; D'Urso, B.; and Gabrielse, G. (2006). "New Measurement of the Electron Magnetic Moment Using a One-Electron Quantum Cyclotron". Physical Review Letters 97 (3): 030801. Bibcode2006PhRvL..97c0801O. doi:10.1103/PhysRevLett.97.030801. PMID16907490. [2] Strictly speaking, "a" refers to the hyperfine splitting constant, a line spacing measured in magnetic field units, while A and B refer to hyperfine coupling constants measured in frequency units. Splitting and coupling constants are proportional, but not identical. The book by Wertz and Bolton has more information (pp. 46 and 442). Wertz, J. E., & Bolton, J. R. (1972). Electron spin resonance: Elementary theory and practical applications. New York: McGraw-Hill. [3] Arthur Schweiger, Gunnar Jeschke (2001). Principles of Pulse Electron Paramagnetic Resonance. Oxford University Press. ISBN978-0198506348. [4] "Dosimetry Systems". Journal of the ICRU 8 (5): 29. 2008. doi:10.1093/jicru/ndn027. [5] Gualtieri, G.; Colacicchia, S, Sgattonic, R., Giannonic, M. (2001). "The Chernobyl Accident: EPR Dosimetry on Dental Enamel of Children". Applied Radiation and Isotopes 55 (1): 7179. doi:10.1016/S0969-8043(00)00351-1. PMID11339534. [6] Chumak, V.; Sholom, S.; Pasalskaya, L. (1999). "Application of High Precision EPR Dosimetry with Teeth for Reconstruction of Doses to Chernobyl Populations" (http:/ / rpd. oxfordjournals. org/ cgi/ content/ abstract/ 84/ 1-4/ 515). Radiation Protection Dosimetry 84: 515520. . [7] Baron, M., Chamayou, A., Marchioro, L., Raffi, J. (2005). "Radicalar probes to measure the action of energy on granular materials". Adv. Powder Technol 16 (3): 199212. doi:10.1163/1568552053750242. [8] EPR of low-dimensional systems (http:/ / hf-epr. sitesled. com)

External links
Electron Magnetic Resonance Program (http://www.magnet.fsu.edu/usershub/scientificdivisions/emr/ overview.html) National High Magnetic Field Laboratory Electron Paramagnetic Resonance (Specialist Periodical Reports) (http://www.rsc.org/shop/books/series. asp?seriesid=49) Published by the Royal Society of Chemistry

Larmor precession

150

Larmor precession
In physics, Larmor precession (named after Joseph Larmor) is the precession of the magnetic moments of electrons, atomic nuclei, and atoms about an external magnetic field. The magnetic field exerts a torque on the magnetic moment,

where

is the torque,

is the magnetic dipole moment, is the external magnetic field,

is the angular symbolizes the

momentum vector, cross product, and

is the gyromagnetic ratio which gives the

proportionality constant between the magnetic moment and the angular momentum.

Larmor frequency
The angular momentum vector precesses about the external field axis with an angular frequency known as the Larmor frequency,

where

is the angular frequency,[1]

is the gyromagnetic ratio, and

is the magnitude of the magnetic

field[2] and is the g-factor (normally 1, except for in quantum physics). Simplified, this becomes:

where is the Larmor frequency, m is mass, e is charge, and B is applied field. For a given nucleus, the g-factor includes the effects of the spin of the nucleons as well as their orbital angular momentum and the coupling between the two. Because the nucleus is so complicated, g factors are very difficult to calculate, but they have been measured to high precision for most nuclei. Each nuclear isotope has a unique Larmor frequency for NMR spectroscopy, which is tabulated here [3].

Including Thomas precession


The above equation is the one that is used in most applications. However, a full treatment must include the effects of Thomas precession, yielding the equation (in CGS units):

where

is the relativistic gamma factor (not to be confused with the gyromagnetic ratio above). Notably, for the

electron g is very close to 2 (2.002..), so if one sets g=2, one arrives at

Larmor precession

151

Bargmann-Michel-Telegdi equation
The spin precession of an electron in an external electromagnetic field is described by the Bargmann-Michel-Telegdi (BMT) equation [4]

where electron,

, and

are polarization four-vector, charge, mass, and magnetic moment, , , and

is four-velocity of

is electromagnetic field-strength tensor. Using

equations of motion,

one can rewrite the first term in the right side of the BMT equation as

, where

is four-acceleration. This term describes Fermi-Walker transport and leads to Thomas precession. The second term is associated with Larmor precession.

Applications
A 1935 paper published by Lev Landau and Evgeny Lifshitz predicted the existence of ferromagnetic resonance of the Larmor precession, which was independently verified in experiments by J. H. E. Griffiths (UK) and E. K. Zavoiskij (USSR) in 1946. Larmor precession is important in nuclear magnetic resonance, electron paramagnetic resonance and muon spin resonance. To calculate the spin of a particle in a magnetic field, one must also take into account Thomas precession.

Notes
[1] Spin Dynamics, Malcolm H. Levitt, Wiley, 2001 [2] Louis N. Hand and Janet D. Finch. (1998). Analytical mechanics (http:/ / books. google. com/ ?id=1J2hzvX2Xh8C& pg=PA192& lpg=PA192& dq=Larmor's+ Theorem). Cambridge, England: Cambridge University Press. p.192. ISBN9780521575720. . [3] http:/ / www-lcs. ensicaen. fr/ pyPulsar/ index. php/ List_of_NMR_isotopes [4] V. Bargmann, L. Michel, and V. L. Telegdi, Precession of the Polarization of Particles Moving in a Homogeneous Electromagnetic Field, Phys. Rev. Lett. 2, 435 (1959).

External links
Georgia State University HyperPhysics page on Larmor Frequency (http://hyperphysics.phy-astr.gsu.edu/ hbase/nuclear/larmor.html) Larmor Frequency Calculator (http://bio.groups.et.byu.net/LarmourFreqCal.phtml)

Pulsed EPR

152

Pulsed EPR
Pulsed electron paramagnetic resonance (EPR) is a spectroscopic technique related to common nuclear magnetic resonance (NMR). Its most basic form involves the alignment of the net magnetization vector of the electron spins in a constant magnetic field. This alignment is perturbed by applying a short oscillating field, usually a microwave pulse. One can then measure the emitted microwave signal which is created by the sample magnetization. Fourier transformation of the microwave signal yields an EPR spectrum in the frequency domain. With a vast variety of pulse sequences Spin echo animation showing the response of electron spins (red arrows) in the blue Bloch sphere to the green pulse sequence it is possible to gain extensive knowledge on structural and dynamical properties of paramagnetic compounds. Pulsed EPR techniques such as electron spin echo envelope modulation (ESEEM) or pulsed electron nuclear double resonance (ENDOR) can reveal the interactions of the electron spin with its surrounding nuclear spins.

Scope
Electron paramagnetic resonance (EPR) or electron spin resonance (ESR) is a spectroscopic technique widely used in biology, chemistry, medicine and physics to study systems with one or more unpaired electrons. Because of the specific relation between the magnetic parameters, electronic wavefunction and the configuration of the surrounding non-zero spin nuclei, EPR and ENDOR provide information on the structure, dynamics and the spatial distribution of the paramagnetic species. However, these techniques are limited in spectral and time resolution when used with traditional continuous wave methods. This resolution can be improved in pulsed EPR by investigating interactions separately from each other via pulse sequences.

Historical overview
R. J. Blume reported the first electron spin echo in 1958, which came from a solution of sodium in ammonia at room temperature [1]. A magnetic field of 0.62 mT was used requiring a frequency of 17.4 MHz. The first microwave electron spin echoes were reported in the same year by Gordon and Bowers using 23 GHz excitation of dopants in silicon[2]. A lot of the pioneering early pulsed EPR was conducted in the group of W. B. Mims at Bell Labs during the 1960s. In the first decade only a small number of groups worked the field, because of the expensive instrumentation, the lack of suitable microwave components and slow digital electronics. The first observation of electron spin echo envelope modulation (ESEEM) was made in 1961 by Mims, Nassau and McGee [3]. Pulsed electron nuclear double resonance (ENDOR) was invented in 1965 by Mims [4]. In this experiment, pulsed NMR transitions are detected with pulsed EPR. ESEEM and pulsed ENDOR continue to be important for studying nuclear spins coupled to electron spins.

Pulsed EPR In the 1980s, the upcoming of the first commercial pulsed EPR and ENDOR spectrometers in the X band frequency range, lead to a fast growth of the field. In the 1990s, parallel to the upcoming high-field EPR, pulsed EPR and ENDOR became a new fast advancing magnetic resonance spectroscopy tool and the first commercial pulsed EPR and ENDOR spectrometer at W band frequencies appeared on the market.

153

Principle
The basic principle of pulsed EPR is similar to NMR spectroscopy. Differences can be found in the relative size of the magnetic interactions and in the relaxation rates which are orders of magnitudes larger in EPR than NMR. A full description of the theory is given within the quantum mechanical formalism, but since the magnetization is being measured as a bulk property, a more intuitive picture can be obtained with a classical description. For a better understanding of the concept of pulsed EPR let us consider the effects on the magnetization vector in the laboratory frame as well as in the rotating frame. In the laboratory frame the static magnetic field B0 is assumed to be parallel to the z-axis and the microwave field B1 parallel to the x-axis. When an electron spin is placed in magnetic field it experiences a torque which causes its magnetic moment to precess around the magnetic field. The precession frequency is known as the Larmor frequency L (see page 18 of ref [5]). where is the gyromagnetic ratio and B0 the magnetic field. The electron spins are characterized by two quantum mechanical states, one parallel and one antiparallel to B0. Because of the lower energy of the parallel state more electron spins can be found in this state according to the Boltzmann distribution. This results in a net magnetization, which is the vector sum of all magnetic moments in the sample, parallel to the z-axis and the magnetic field. To better comprehend the effects of the microwave field B1 it is easier to move to the rotating frame. EPR experiments usually use a microwave resonator designed to create a linearly polarized microwave field B1, perpendicular to the much stronger applied magnetic field B0. The rotating frame is fixed to the rotating B1 components. First we assume to be on resonance with the precessing magnetization vector M0. Therefore the component of B1 will appear stationary. In this frame also the precessing magnetization components appear to be stationary that leads to the disappearance of B0, and we need only to consider B1 and M0. The M0 vector is under the influence of the stationary field B1, leading to another precession of M0, this time around B1 at the frequency 1. This angular frequency 1 is also called the Rabi frequency. Assuming B1 to be parallel to the x-axis, the magnetization vector will rotate around the +x-axis in the zy-plane as long as the microwaves are applied. The angle by which M0 is rotated is called the tip angle and is given by: Here tp is the duration for which B1 is applied, also called the pulse length. The pulses are labeled by the rotation of M0 which they cause and the direction from which they are coming from, since the microwaves can be phase-shifted from the x-axis on to the y-axis. For example, a +y /2 pulse means that a B1 field, which has been 90 degrees phase-shifted out of the +x into the +y direction, has rotated M0 by a tip angle of /2, hence the magnetization would end up along the x-axis. That means the end position of the magnetization vector M0 depends on the length, the magnitude and direction of the microwave pulse B1. In order to understand how the sample emits microwaves after the intense microwave pulse we need to go back to the laboratory frame. In the rotating frame and on resonance the magnetization appeared to be stationary along the x or y-axis after the pulse. In the laboratory frame it becomes a rotating magnetization in the x-y plane at the Larmor frequency. This rotation generates a signal which is maximized if the magnetization vector is exactly in the xy-plane. This microwave signal generated by the rotating magnetization vector is called free induction decay (FID) (see page 175 of ref [5]).

Pulsed EPR Another assumption we have made was the exact resonance condition, in which the Larmor frequency is equal to the microwave frequency. In reality EPR spectra have many different frequencies and not all of them can be exactly on resonance, therefore we need to take off-resonance effects into account. The off-resonance effects lead to three main consequences. The first consequence can be better understood in the rotating frame. A /2 pulse leaves magnetization in the xy-plane, but since the microwave field (and therefore the rotating frame) do not have the same frequency as the precessing magnetization vector, the magnetization vector rotates in the xy-plane, either faster or slower than the microwave magnetic field B1. The rotation rate is governed by the frequency difference . If is 0 then the microwave field rotates as fast as the magnetization vector and both appear to be stationary to each other. If >0 then the magnetization rotates faster than the microwave field component in a counter-clockwise motion and if <0 then the magnetization is slower and rotates clockwise. This means that the individual frequency components of the EPR spectrum, will appear as magnetization components rotating in the xy-plane with the rotation frequency . The second consequence appears in the laboratory frame. Here B1 tips the magnetization differently out of the z-axis, since B0 does not disappear when not on resonance due to the precession of the magnetization vector at . That means that the magnetization is now tipped by an effective magnetic field Beff, which originates from the vector sum of B1 and B0. The magnetization is then tipped around Beff at a faster effective rate eff. This leads directly to the third consequence that the magnetization can not be efficiently tipped into the xy-plane because Beff does not lie in the xy-plane, as B1 does. The motion of the magnetization now defines a cone. That means as becomes larger, the magnetization is tipped less effectively into the xy-plane, and the FID signal decreases. In broad EPR spectra where > 1 it is not possible to tip all the magnetization into the xy-plane to generate a strong FID signal. This is why it is important to maximize 1 or minimize the /2 pulse length for broad EPR signals. So far the magnetization was tipped into the xy-plane and it remained there with the same magnitude. However, in reality the electron spins interact with their surroundings and the magnetization in the xy-plane will decay and eventually return to alignment with the z-axis. This relaxation process is described by the spin-lattice relaxation time T1, which is a characteristic time needed by the magnetization to return to the z-axis, and by the spin-spin relaxation time T2, which describes the vanishing time of the magnetization in the xy-plane. The spin-lattice relaxation results from the urge of the system to return to thermal equilibrium after it has been perturbed by the B1 pulse. Return of the magnetization parallel to B0 is achieved through interactions with the surroundings, that is spin-lattice relaxation. The corresponding relaxation time needs to be considered when extracting a signal from noise, where the experiment needs to be repeated several times, as quickly as possible. In order to repeat the experiment, one needs to wait until the magnetization along the z-axis has recovered, because if there is no magnetization in z direction, then there is nothing to tip into the xy-plane to create a significant signal. The spin-spin relaxation time, also called the transverse relaxation time, is related to homogeneous and inhomogeneous broadening. An inhomogeneous broadening results from the fact that the different spins experience local magnetic field inhomogeneities (different surroundings) creating a large number of spin packets characterized by a distribution of . As the net magnetization vector precesses, some spin packets slow down due to lower fields and others speed up due to higher fields leading to a fanning out of the magnetization vector that results in the decay of the EPR signal. The other packets contribute to the transverse magnetization decay due to the homogeneous broadening. In this process all the spin in one spin packet experience the same magnetic field and interact with each other that can lead to mutual and random spin flip-flops. These fluctuations contribute to a faster fanning out of the magnetization vector. All the information about the frequency spectrum is encoded in the motion of the transverse magnetization. The frequency spectrum is reconstructed using the time behavior of the transverse magnetization made up of y- and x-axis components. It is convenient that these two can be treated as the real and imaginary components of a complex

154

Pulsed EPR quantity and use the Fourier theory to transform the measured time domain signal into the frequency domain representation. This is possible because both the absorption (real) and the dispersion (imaginary) signals are detected. The FID signal decays away and for very broad EPR spectra this decay is rather fast due to the inhomogeneous broadening. To obtain more information one can recover the disappeared signal with another microwave pulse to produce a Hahn echo[6]. After applying a /2 pulse (90), the magnetization vector is tipped into the xy-plane producing an FID signal. Different frequencies in the EPR spectrum (inhomogeneous broadening) cause this signal to "fan out", meaning that the slower spin-packets trail behind the faster ones. After a certain time t, a pulse (180) is applied to the system inverting the magnetization, and the fast spin-packets are then behind catching up with the slow spin-packets. A complete refocusing of the signal occurs then at time 2t. An accurate echo caused by a second microwave pulse can remove all inhomogeneous broadening effects. After all of the spin-packets bunch up, they will dephase again just like an FID. In other words, a spin echo is a reversed FID followed by a normal FID, which can be Fourier transformed to obtain the EPR spectrum. The longer the time between the pulses becomes, the smaller the echo will be due to spin relaxation. When this relaxation leads to an exponential decay in the echo height, the decay constant is the phase memory time TM, which can have many contributions such as transverse relaxation, spectral, spin and instantaneous diffusion. Changing the times between the pulses leads to a direct measurement of TM as shown in the spin echo decay animation below.

155

Applications
ESEEM [3] [5] and pulsed ENDOR [4] [5] are widely-used echo experiments, in which the interaction of electron spins with the nuclei in their environment can be studied and controlled. Quantum computing and spintronics, in which spins are used to store information, have led to new lines of research in pulsed EPR. One of the most popular pulsed EPR experiments currently is double electron-electron resonance (DEER), which is also known as pulsed electron-electron double resonance (ELDOR). [5] This uses two different frequencies to control different spins in order to find out the strength of their coupling. The distance between the spins can then be inferred from their coupling strength, which is used to study structures of large bio-molecules.

References
[1] Blume, R. J. (1958). "Electron Spin Relaxation Times in Sodium-Ammonia Solutions". Physical Review 109: 18671873. Bibcode1958PhRv..109.1867B. doi:10.1103/PhysRev.109.1867. [2] Gordon, J. P.; Bowers, K. D. (1958). "Microwave Spin Echoes from Donor Electrons in Silicon". Physical Review Letters 1: 368370. Bibcode1958PhRvL...1..368G. doi:10.1103/PhysRevLett.1.368. [3] Mims, W. B.; Nassau, K.; McGee J. D. (1961). "Spectral Diffusion in Electron Resonance Lines". Physical Review 123: 20592069. Bibcode1961PhRv..123.2059M. doi:10.1103/PhysRev.123.2059. [4] Mims, W. B. (1965). "Pulsed endor experiments". Proceedings of the Royal Society of London Series A-Mathematical and Physical Sciences 283: 452457. Bibcode1965RSPSA.283..452M. doi:10.1098/rspa.1965.0034. [5] Schweiger, A. and Jeschke, G. (2001). Principles of pulse electron paramagnetic resonance (http:/ / books. google. com/ books?id=tkvQQElkW1wC& printsec=frontcover). Oxford University Press, New York. ISBN0198506341. . [6] Hahn, E.L. (1950). "Spin echoes". Physical Review 80: 580594. Bibcode1950PhRv...80..580H. doi:10.1103/PhysRev.80.580.

External links
Pulse EPR (http://www.bruker-biospin.com/pulse.html) Bruker BioSpin

156

Quantum Magnetism
Bohr magneton
The value of Bohr magneton
system of units SI [1] value 9.27400915(23)1024 9.27400915(23)1021 unit JT1 ErgG1

[2] CGS eV [3]

5.7883817555(79)105 eVT1
1

atomic units

none

In atomic physics, the Bohr magneton (symbol B) is a physical constant and the natural unit for expressing an electron magnetic dipole moment. The Bohr magneton is defined in SI units by

and in Gaussian CGS units by

where e is the elementary charge, is the reduced Planck constant, me is the electron rest mass and c is the speed of light. The magnitude of an electron's spin magnetic moment is approximately one Bohr magneton.[4]

History
The idea of elementary magnets is due to Walter Ritz (1907) and Pierre Weiss. Already before the Rutherford model of atomic structure, several theorists commented that the magneton should involve Planck's constant h.[5] By postulating that the ratio of electron kinetic energy to orbital frequency should be equal to h, Richard Gans computed a value that was twice as large as the Bohr magneton in September 1911.[6] At the First Solvay Conference in November that year, Paul Langevin obtained a submultiple.[7] The Romanian physicist tefan Procopiu obtained for the first time its value in 1911;[8][9] the value is referred to as the "BohrProcopiu magneton" in Romanian scientific literature.[10] The Bohr magneton is the magnitude of the magnetic dipole moment of an orbiting electron with an orbital angular momentum of one . According to the Bohr model, this is the ground state, i.e. the state of lowest possible energy.[11] In the summer of 1913, this value was naturally obtained by the Danish physicist Niels Bohr as a consequence of his atom model,[6][12] and also published independently by Procopiu using directly Max Planck's quantum theory.[9] In 1920, Wolfgang Pauli gave the Bohr magneton its name in an article where he contrasted it with the magneton of the experimentalists which he called the Weiss magneton.[5]

Bohr magneton

157

References
[1] "CODATA value: Bohr magneton" (http:/ / physics. nist. gov/ cgi-bin/ cuu/ Value?mub). The NIST Reference on Constants, Units, and Uncertainty. NIST. . Retrieved 2009-12-22. [2] Robert C. O'Handley (2000). Modern magnetic materials: principles and applications. John Wiley & Sons. p.83. ISBN0-471-15566-7. [3] "CODATA value: Bohr magneton in eV/T" (http:/ / physics. nist. gov/ cgi-bin/ cuu/ Value?mubev). The NIST Reference on Constants, Units, and Uncertainty. NIST. . Retrieved 2010-08-14. [4] Anant S. Mahajan, Abbas A. Rangwala (1989). Electricity and Magnetism (http:/ / books. google. com/ ?id=_tXrjggX7WwC& pg=PA419& lpg=PA419& dq="intrinsic+ dipole+ moment"+ and+ electron+ "Bohr+ magneton"). McGraw-Hill. p.419. ISBN978-0-07-460225-6. . [5] Stephen T. Keith and Pierre Qudec (1992). "Magnetism and Magnetic Materials: The Magneton". Out of the Crystal Maze. pp.384394. ISBN0-19-505329. [6] John Heilbron; Thomas Kuhn (1969). "The genesis of the Bohr atom". Historical Studies in the Physical Sciences 1: 232. [7] Paul Langevin (1911). "La thorie cintique du magntisme et les magntons". La thorie du rayonnement et les quanta: Rapports et discussions de la runion tenue Bruxelles, du 30 octobre au 3 novembre 1911, sous les auspices de M. E. Solvay (http:/ / www. archive. org/ details/ lathoriedurayo00inst). p.403. [8] tefan Procopiu (19111913). "Sur les lments dnergie". Annales scientifiques de l'Universit de Jassy 7: 280. [9] tefan Procopiu (1913). "Determining the Molecular Magnetic Moment by M. Planck's Quantum Theory". Bulletin scientifique de lAcadmie roumaine de sciences 1: 151. [10] "Stefan Procopiu (1890-1972)" (http:/ / www. etti. tuiasi. ro/ sibm/ old/ Technical Museum/ html/ en/ Stefan_Procopiu_en. htm). Stefan Procopiu Science and Technique Museum. . Retrieved 2010-11-03. [11] Marcelo Alonso, Edward Finn (1992). Physics. Addison-Wesley. ISBN978-0-201-56518-8. [12] Abraham Pais (1991). Niels Bohr's Times, in physics, philosophy, and politics. Clarendon Press. ISBN0-19-852048-4.

Spin (physics)
In quantum mechanics and particle physics, spin is a fundamental characteristic property of elementary particles, composite particles (hadrons), and atomic nuclei.[1] All elementary particles of a given kind have the same spin quantum number, an important part of the quantum state of a particle. When combined with the spin-statistics theorem, the spin of electrons results in the Pauli exclusion principle, which in turn underlies the periodic table of chemical elements. The spin direction (also called spin for short) of a particle is an important intrinsic degree of freedom. Wolfgang Pauli was the first to propose the concept of spin, but he did not name it. In 1925, Ralph Kronig, George Uhlenbeck, and Samuel Goudsmit suggested a physical interpretation of particles spinning around their own axis. The mathematical theory was worked out in depth by Pauli in 1927. When Paul Dirac derived his relativistic quantum mechanics in 1928, electron spin was an essential part of it. There are two types of angular momentum in quantum mechanics: Orbital angular momentum, which is a generalization of angular momentum in classical mechanics (L=rp), and spin, which has no analogue in classical mechanics.[2][3] Since spin is a type of angular momentum, it has the same dimensions: Js in SI units. In practice, however, SI units are never used to describe spin: instead, it is written as a multiple of the reduced Planck constant . In natural units, the is omitted, so spin is written as a unitless number. The spin quantum numbers are always unitless numbers by definition.

Spin (physics)

158

Spin quantum number


As the name suggests, spin was originally conceived as the rotation of a particle around some axis. This picture is correct so far as spins obey the same mathematical laws as quantized angular momenta do. On the other hand, spins have some peculiar properties that distinguish them from orbital angular momenta: Spin quantum numbers may take on with a gluon (the green ball) from another proton with opposite spin; spin is half-odd-integer values; represented by the blue arrows circling the protons and the quark. The blue Although the direction of its spin can be question marks circling the gluon represent the question: Are gluons polarised? The particles ejected from the collision are a shower of quarks and one photon of changed, an elementary particle cannot light (the purple ball). be made to spin faster or slower. The spin of a charged particle is associated with a magnetic dipole moment with a g-factor differing from 1. This could only occur classically if the internal charge of the particle were distributed differently from its mass.
The head-on collision of a quark (the red ball) from one proton (the orange ball)

Elementary particles
Elementary particles are particles for which there is no known method of division into smaller units. Theoretical and experimental studies have shown that the spin possessed by such particles cannot be explained by postulating that they are made up of even smaller particles rotating about a common center of mass (see classical electron radius); as far as can be determined, these elementary particles have no inner structure. The spin of an elementary particle is a truly intrinsic physical property, akin to the particle's electric charge and rest mass. The conventional definition of the spin quantum number s is s=n/2, where n can be any non-negative integer. Hence the allowed values of s are 0, 1/2, 1, 3/2, 2, etc. The value of s for an elementary particle depends only on the type of particle, and cannot be altered in any known way (in contrast to the spin direction described below). The spin angular momentum S of any physical system is quantised. The allowed values of S are:

Father and Mother of the series Spin Family (2009) by physicist-turned-sculptor Julian Voss-Andreae. The two pictured objects illustrate the geometry of a spin 5/2 object (blue 'male' on the left) and spin 2 object (pink 'female' on the right). Spin Family, on display in the art exhibition "Quantum Objects" playfully equates fermions with the male and bosons with the female gender, depicting the first spin 1/2, 1, 3/2, [4] 2, and 5/2 objects as a family of five.

where h is the Planck constant. In contrast, orbital angular momentum can only take on integer values of s, even values of n. That is why rather than was defined as the quantum mechanical unit of angular

momentum. When spin was discovered it was too late to change. All known matter is ultimately composed of elementary particles called fermions, and all elementary fermions have s=1/2. Examples of fermions are the electron and positron, the quarks making up protons

and neutrons, and the neutrinos. Elementary particles emit and receive one or more particles called bosons. This boson exchange gives rise to the three fundamental interactions ("forces") of the Standard model of particle physics;

Spin (physics) hence bosons are also called force carriers. These bosons have s=1. A basic example of a boson is the photon. Electromagnetism is the force that results when charged particles exchange photons. Theory predicts the existence of two bosons whose s differs from 1. The force carrier for gravity is the hypothetical graviton; theory suggests that it has s=2. The Higgs mechanism predicts that elementary particles acquire nonzero rest mass by exchanging hypothetical Higgs bosons with an all-pervasive Higgs field. Theory predicts that the Higgs boson has s=0. If so, it would be the only elementary particle for which this is the case.

159

Composite particles
The spin of composite particles, such as protons, neutrons, and atomic nuclei is usually understood to mean the total angular momentum. This is the sum of the spins and orbital angular momenta of the constituent particles. Such a composite spin is subject to the same quantization condition as any other angular momentum. Composite particles are often referred to as having a definite spin, just like elementary particles; for example, the proton is a spin-1/2 particle. This is understood to refer to the spin of the lowest-energy internal state of the composite particle (i.e., a given spin and orbital configuration of the constituents).[5] It is not always easy to deduce the spin of a composite particle from first principles. For example, even though we know that the proton is a spin-1/2 particle, the question of how this spin is distributed among the three internal valence quarks and the surrounding sea quarks and gluons is an active area of research. Delta baryons, which decay into protons and neutrons, have spin 3/2. All the three quarks inside a Delta baryon () have their spin axis pointing in the same direction, unlike the nearly identical proton and neutron (called "nucleons") in which the intrinsic spin of one of the three constituent quarks is always opposite the spin of the other two. This difference in spin alignment is the only quantum number distinction between the + and 0 and ordinary nucleons.

Atoms and molecules


The spin of atoms and molecules is the sum of the spins of unpaired electrons, which may be parallel or antiparallel. It is responsible for paramagnetism.

The spin-statistics theorem


The spin of a particle has crucial consequences for its properties in statistical mechanics. Particles with half-integer spin obey Fermi-Dirac statistics, and are known as fermions. They are required to occupy antisymmetric quantum states (see the article on identical particles.) This property forbids fermions from sharing quantum states a restriction known as the Pauli exclusion principle. Particles with integer spin, on the other hand, obey Bose-Einstein statistics, and are known as bosons. These particles occupy "symmetric states", and can therefore share quantum states. The proof of this is known as the spin-statistics theorem, which relies on both quantum mechanics and the theory of special relativity. In fact, "the connection between spin and statistics is one of the most important applications of the special relativity theory".[6]

Spin (physics)

160

Magnetic moments
Particles with spin can possess a magnetic dipole moment, just like a rotating electrically charged body in classical electrodynamics. These magnetic moments can be experimentally observed in several ways, e.g. by the deflection of particles by inhomogeneous magnetic fields in a SternGerlach experiment, or by measuring the magnetic fields generated by the particles themselves. The intrinsic magnetic moment of a Spin- particle with charge q, mass m, and spin angular momentum S, is[7]

where the dimensionless quantity gs is called the spin g-factor. For exclusively orbital rotations it would be 1 (assuming that the mass and the charge occupy spheres of equal radius).

Magnetic field lines around a magnetostatic dipole; the magnetic dipole itself is in the center and is seen from the side.

The electron, being a charged elementary particle, possesses a nonzero magnetic moment. One of the triumphs of the theory of quantum electrodynamics is its accurate prediction of the electron g-factor, which has been experimentally determined to have the value 2.0023193043622(15), with the digits in parentheses denoting measurement uncertainty in the last two digits at one standard deviation.[8] The value of 2 arises from the Dirac equation, a fundamental equation connecting the electron's spin with its electromagnetic properties, and the correction of 0.002319304... arises from the electron's interaction with the surrounding electromagnetic field, including its own field.[9] Composite particles also possess magnetic moments associated with their spin. In particular, the neutron possesses a non-zero magnetic moment despite being electrically neutral. This fact was an early indication that the neutron is not an elementary particle. In fact, it is made up of quarks, which are electrically charged particles. The magnetic moment of the neutron comes from the spins of the individual quarks and their orbital motions. Neutrinos are both elementary and electrically neutral. The minimally extended Standard Model that takes into account non-zero neutrino masses predicts neutrino magnetic moments of:[10][11][12]

where the are the neutrino magnetic moments, m are the neutrino masses, and B is the Bohr magneton. New physics above the electroweak scale could, however, lead to significantly higher neutrino magnetic moments. It can be shown in a model independent way that neutrino magnetic moments larger than about 1014B are unnatural, because they would also lead to large radiative contributions to the neutrino mass. Since the neutrino masses cannot exceed about 1eV, these radiative corrections must then be assumed to be fine tuned to cancel out to a large degree.[13] The measurement of neutrino magnetic moments is an active area of research. As of 2001, the latest experimental results have put the neutrino magnetic moment at less than 1.21010 times the electron's magnetic moment. In ordinary materials, the magnetic dipole moments of individual atoms produce magnetic fields that cancel one another, because each dipole points in a random direction. Ferromagnetic materials below their Curie temperature, however, exhibit magnetic domains in which the atomic dipole moments are locally aligned, producing a macroscopic, non-zero magnetic field from the domain. These are the ordinary "magnets" with which we are all familiar. In paramagnetic materials, the magnetic dipole moments of individual atoms spontaneously align with an externally applied magnetic field. In diamagnetic materials, on the other hand, the magnetic dipole moments of individual atoms spontaneously align oppositely to any externally applied magnetic field, even if it requires energy to do so.

Spin (physics) The study of the behavior of such "spin models" is a thriving area of research in condensed matter physics. For instance, the Ising model describes spins (dipoles) that have only two possible states, up and down, whereas in the Heisenberg model the spin vector is allowed to point in any direction. These models have many interesting properties, which have led to interesting results in the theory of phase transitions.

161

Spin direction
Spin projection quantum number and spin multiplicity
In classical mechanics, the angular momentum of a particle possesses not only a magnitude (how fast the body is rotating), but also a direction (either up or down on the axis of rotation of the particle). Quantum mechanical spin also contains information about direction, but in a more subtle form. Quantum mechanics states that the component of angular momentum measured along any direction can only take on the values [14]

where Si is the spin component along the i-axis (either x, y, or z), si is the spin projection quantum number along the i-axis, and s is the principal spin quantum number (discussed in the previous section). Conventionally the direction chosen is the z-axis:

where Sz is the spin component along the z-axis, sz is the spin projection quantum number along the z-axis. One can see that there are 2s+1 possible values of sz. The number "2s+1" is the multiplicity of the spin system. For example, there are only two possible values for a spin-1/2 particle: sz=+1/2 and sz=1/2. These correspond to quantum states in which the spin is pointing in the +z or z directions respectively, and are often referred to as "spin up" and "spin down". For a spin-3/2 particle, like a delta baryon, the possible values are +3/2, +1/2, 1/2, 3/2.

Spin vector
For a given quantum state, one could think of a spin vector the spin components along each axis, i.e., whose components are the expectation values of . This vector then would describe the

"direction" in which the spin is pointing, corresponding to the classical concept of the axis of rotation. It turns out that the spin vector is not very useful in actual quantum mechanical calculations, because it cannot be measured directly sx, sy and sz cannot possess simultaneous definite values, because of a quantum uncertainty relation between them. However, for statistically large collections of particles that have been placed in the same pure quantum state, such as through the use of a Stern-Gerlach apparatus, the spin vector does have a well-defined experimental meaning: It specifies the direction in ordinary space in which a subsequent detector must be oriented in order to achieve the maximum possible probability (100%) of detecting every particle in the collection. For spin-1/2 particles, this maximum probability drops off smoothly as the angle between the spin vector and the detector increases, until at an angle of 180 degreesthat is, for detectors oriented in the opposite direction to the spin vectorthe expectation of detecting particles from the collection reaches a minimum of 0%. As a qualitative concept, the spin vector is often handy because it is easy to picture classically. For instance, quantum mechanical spin can exhibit phenomena analogous to classical gyroscopic effects. For example, one can exert a kind of "torque" on an electron by putting it in a magnetic field (the field acts upon the electron's intrinsic magnetic dipole momentsee the following section). The result is that the spin vector undergoes precession, just like a classical gyroscope. This phenomenon is used in nuclear magnetic resonance sensing. Mathematically, quantum mechanical spin is not described by vectors as in classical angular momentum, but by objects known as spinors. There are subtle differences between the behavior of spinors and vectors under coordinate rotations. For example, rotating a spin-1/2 particle by 360 degrees does not bring it back to the same quantum state, but to the state with the opposite quantum phase; this is detectable, in principle, with interference experiments. To

Spin (physics) return the particle to its exact original state, one needs a 720 degree rotation. A spin-zero particle can only have a single quantum state, even after torque is applied. Rotating a spin-2 particle 180 degrees can bring it back to the same quantum state and a spin-4 particle should be rotated 90 degrees to bring it back to the same quantum state. The spin 2 particle can be analogous to a straight stick that looks the same even after it is rotated 180 degrees and a spin 0 particle can be imagined as sphere which looks the same after whatever angle it is turned through.

162

Mathematical formulation of spin


Spin operator
Spin obeys commutation relations analogous to those of the orbital angular momentum: is the Levi-Civita symbol. It follows (as with angular momentum) that the eigenvectors of S2 and Sz

where

(expressed as kets in the total S basis) are:

The spin raising and lowering operators acting on these eigenvectors give: , where But unlike orbital angular momentum the eigenvectors are not spherical harmonics. They are not functions of and . There is also no reason to exclude half-integer values of s and m. In addition to their other properties, all quantum mechanical particles possess an intrinsic spin (though it may have the intrinsic spin 0, too). The spin is quantized in units of the reduced Planck constant, such that the state function of the particle is, say, not , but where is out of the following discrete set of values:

One distinguishes bosons (integer spin) and fermions (half-integer spin). The total angular momentum conserved in interaction processes is then the sum of the orbital angular momentum and the spin.

Pauli matrices and spin operators


The quantum mechanical operators associated with spin observables are:

In the special case of spin-1/2 particles, x, y and z are the three Pauli matrices, given by:

Spin and the Pauli exclusion principle


For systems of N identical particles this is related to the Pauli exclusion principle, which states that by interchanges of any two of the N particles one must have Thus, for bosons the prefactor (1)2s will reduce to +1, for fermions to 1. In quantum mechanics all particles are either bosons or fermions. In some speculative relativistic quantum field theories "supersymmetric" particles also exist, where linear combinations of bosonic and fermionic components appear. In two dimensions, the prefactor (1)2s can be replaced by any complex number of magnitude 1 (see Anyon).

Spin (physics) Electrons are fermions with s=1/2; quanta of light ("photons") are bosons with s=1. This shows also explicitly that the property spin cannot be fully explained as a classical intrinsic orbital angular momentum, e.g., similar to that of a "spinning top", since orbital angular rotations would lead to integer values of s. Instead one is dealing with an essential legacy of relativity. The photon, in contrast, is always relativistic (velocity vc), and the corresponding classical theory, that of Maxwell, is also relativistic. The above permutation postulate for N-particle state functions has most-important consequences in daily life, e.g. the periodic table of the chemists or biologists.

163

Spin and rotations


As described above, quantum mechanics states that component of angular momentum measured along any direction can only take a number of discrete values. The most convenient quantum mechanical description of particle's spin is therefore with a set of complex numbers corresponding to amplitudes of finding a given value of projection of its intrinsic angular momentum on a given axis. For instance, for a spin 1/2 particle, we would need two numbers a1/2, giving amplitudes of finding it with projection of angular momentum equal to /2 and /2, satisfying the requirement

For a generic particle with spin s, we would need 2s+1 such parameters. Since these numbers depend on the choice of the axis, they transform into each other non-trivially when this axis is rotated. It's clear that the transformation law must be linear, so we can represent it by associating a matrix with each rotation, and the product of two transformation matrices corresponding to rotations A and B must be equal (up to phase) to the matrix representing rotation AB. Further, rotations preserve the quantum mechanical inner product, and so should our transformation matrices:

Mathematically speaking, these matrices furnish a unitary projective representation of the rotation group SO(3). Each such representation corresponds to a representation of the covering group of SO(3), which is SU(2). There is one n-dimensional irreducible representation of SU(2) for each dimension, though this representation is n-dimensional real for odd n and n-dimensional complex for even n (hence of real dimension 2n). For a rotation by angle in the plane with normal vector , U can be written

where

and

is the vector of spin operators.

A generic rotation in 3-dimensional space can be built by compounding operators of this type using Euler angles:

An irreducible representation of this group of operators is furnished by the Wigner D-matrix:

where

is Wigner's small d-matrix. Note that for = 2 and = = 0, i.e. a full rotation about the z-axis, the Wigner D-matrix elements become

Spin (physics) Recalling that a generic spin state can be written as a superposition of states with definite m, we see that if s is an integer, the values of m are all integers, and this matrix corresponds to the identity operator. However, if s is a half-integer, the values of m are also all half-integers, giving (-1)2m = -1 for all m, and hence upon rotation by 2 the state picks up a minus sign. This fact is a crucial element of the proof of the spin-statistics theorem.

164

Spin and Lorentz transformations


We could try the same approach to determine the behavior of spin under general Lorentz transformations, but we'd immediately discover a major obstacle. Unlike SO(3), the group of Lorentz transformations SO(3,1) is non-compact and therefore does not have any faithful, unitary, finite-dimensional representations. In case of spin 1/2 particles, it is possible to find a construction that includes both a finite-dimensional representation and a scalar product that is preserved by this representation. We associate a 4-component Dirac spinor with each particle. These spinors transform under Lorentz transformations according to the law

where

are gamma matrices and

is an antisymmetric 4x4 matrix parametrizing the transformation. It can be

shown that the scalar product

is preserved. It is not, however, positive definite, so the representation is not unitary.

Measuring spin along the x, y, and z axes


Each of the (Hermitian) Pauli matrices has two eigenvalues, +1 and 1. The corresponding normalized eigenvectors are:

By the postulates of quantum mechanics, an experiment designed to measure the electron spin on the x, y or z axis can only yield an eigenvalue of the corresponding spin operator (Sx, Sy or Sz) on that axis, i.e. /2 or /2. The quantum state of a particle (with respect to spin), can be represented by a two component spinor:

When the spin of this particle is measured with respect to a given axis (in this example, the x-axis), the probability that its spin will be measured as /2 is just . Correspondingly, the probability that its spin will be measured as /2 is just . Following the measurement, the spin state of the particle will collapse into , etc), provided that no

the corresponding eigenstate. As a result, if the particle's spin along a given axis has been measured to have a given eigenvalue, all measurements will yield the same eigenvalue (since measurements of the spin are made along other axes (see compatibility section below).

Spin (physics)

165

Measuring spin along an arbitrary axis


The operator to measure spin along an arbitrary axis direction is easily obtained from the Pauli spin matrices. Let u = (ux, uy, uz) be an arbitrary unit vector. Then the operator for spin in this direction is simply . The operator Su has eigenvalues of /2, just like the usual spin matrices. This method of finding the operator for spin in an arbitrary direction generalizes to higher spin states, one takes the dot product of the direction with a vector of the three operators for the three x, y, z axis directions. A normalized spinor for spin-1/2 in the (ux, uy, uz) direction (which works for all spin states except spin down where it will give 0/0), is:

The above spinor is obtained in the usual way by diagonalizing the matrix and finding the eigenstates corresponding to the eigenvalues. In quantum mechanics, vectors are termed "normalized" when multiplied by a normalizing factor, which results in the vector having a length of unity.

Compatibility of spin measurements


Since the Pauli matrices do not commute, measurements of spin along the different axes are incompatible. This means that if, for example, we know the spin along the x-axis, and we then measure the spin along the y-axis, we have invalidated our previous knowledge of the x-axis spin. This can be seen from the property of the eigenvectors (i.e. eigenstates) of the Pauli matrices that:

So when physicists measure the spin of a particle along the x-axis as, for example, /2, the particle's spin state collapses into the eigenstate . When we then subsequently measure the particle's spin along the y-axis, the spin state will now collapse into either or , each with probability 1/2. Let us say, in our example, and respectively).

that we measure /2. When we now return to measure the particle's spin along the x-axis again, the probabilities that we will measure /2 or /2 are each 1/2 (i.e. they are x-axis will now be measured to have either eigenvalue with equal probability. This implies that the original measurement of the spin along the x-axis is no longer valid, since the spin along the

Spin and parity


In tables of the spin quantum number s for nuclei or particles, the spin is often followed by a "+" or "-". This refers to the parity with "+" for even parity (wave function unchanged by spatial inversion) and "-" for odd parity (wave function negated by spatial inversion). For example, see the isotopes of bismuth.

Applications
Spin has important theoretical implications and practical applications. Well-established direct applications of spin include: Nuclear magnetic resonance spectroscopy in chemistry; Electron spin resonance spectroscopy in chemistry and physics; Magnetic resonance imaging (MRI) in medicine, which relies on proton spin density; Giant magnetoresistive (GMR) drive head technology in modern hard disks.

Spin (physics) Electron spin plays an important role in magnetism, with applications for instance in computer memories. The manipulation of nuclear spin by radiofrequency waves (nuclear magnetic resonance) is important in chemical spectroscopy and medical imaging. Spin-orbit coupling leads to the fine structure of atomic spectra, which is used in atomic clocks and in the modern definition of the second. Precise measurements of the g-factor of the electron have played an important role in the development and verification of quantum electrodynamics. Photon spin is associated with the polarization of light. A possible future direct application of spin is as a binary information carrier in spin transistors. Original concept proposed in 1990 is known as Datta-Das spin transistor.[15] Electronics based on spin transistors is called spintronics, which includes the manipulation of spins in semiconductor devices. There are many indirect applications and manifestations of spin and the associated Pauli exclusion principle, starting with the periodic table of chemistry.

166

History
Spin was first discovered in the context of the emission spectrum of alkali metals. In 1924 Wolfgang Pauli introduced what he called a "two-valued quantum degree of freedom" associated with the electron in the outermost shell. This allowed him to formulate the Pauli exclusion principle, stating that no two electrons can share the same quantum state at the same time. The physical interpretation of Pauli's "degree of freedom" was initially unknown. Ralph Kronig, one of Land's assistants, suggested in early 1925 that it was produced by the self-rotation of the electron. When Pauli heard about the idea, he criticized it severely, noting that the electron's hypothetical surface would have to be moving faster than the speed of light in order for it to rotate quickly enough to produce the necessary angular momentum. This would violate the theory of relativity. Largely due to Pauli's criticism, Kronig decided not to publish his idea. In the autumn of 1925, the same thought came to two Dutch physicists, George Uhlenbeck and Samuel Goudsmit. Under the advice of Paul Ehrenfest, they published their results. It met a favorable response, especially after Llewellyn Thomas managed to resolve a factor-of-two discrepancy between experimental results and Uhlenbeck and Goudsmit's calculations (and Kronig's unpublished ones). This discrepancy was due to the orientation of the electron's tangent frame, in addition to its position.

Wolfgang Pauli

Mathematically speaking, a fiber bundle description is needed. The tangent bundle effect is additive and relativistic; that is, it vanishes if c goes to infinity. It is one half of the value obtained without regard for the tangent space orientation, but with opposite sign. Thus the combined effect differs from the latter by a factor two (Thomas precession). Despite his initial objections, Pauli formalized the theory of spin in 1927, using the modern theory of quantum mechanics discovered by Schrdinger and Heisenberg. He pioneered the use of Pauli matrices as a representation of the spin operators, and introduced a two-component spinor wave-function. Pauli's theory of spin was non-relativistic. However, in 1928, Paul Dirac published the Dirac equation, which described the relativistic electron. In the Dirac equation, a four-component spinor (known as a "Dirac spinor") was used for the electron wave-function. In 1940, Pauli proved the spin-statistics theorem, which states that fermions have half-integer spin and bosons integer spin. In retrospect, the first direct experimental evidence of the electron spin was the Stern-Gerlach experiment of 1922. However, the correct explanation of this experiment was only given in 1927.[16]

Spin (physics)

167

Notes
[1] It is worth noting that the intrinsic property of subatomic particles called spin and discussed in this article, is related in some small ways, but is very different from the everyday concept of spin (for example, as used when describing a spinning ball). Spin, as used by particle physicists in the quantum world, is a property of subatomic particles, which has certain qualities and obeys certain rules. [2] "Angular Momentum Operator Algebra", class notes by Michael Fowler (http:/ / galileo. phys. virginia. edu/ classes/ 751. mf1i. fall02/ AngularMomentum. htm) [3] A modern approach to quantum mechanics, by Townsend, p31 and p80 (http:/ / books. google. com/ books?id=3_7uriPX028C& pg=PA31) [4] Ball, Philip (26 November 2009). "Quantum objects on show" (http:/ / www. nature. com/ nature/ journal/ v462/ n7272/ pdf/ 462416a. pdf). Nature 462 (7272): 416. Bibcode2009Natur.462..416B. doi:10.1038/462416a. . Retrieved 2009-01-12. [5] P. Lemmens and P. Millet (2004). "Spin - Orbit - Topology, a Triptych" (http:/ / www. fkf. mpg. de/ keimer/ Publist/ PDF_2004/ Lemmens_05. pdf). Lect. Notes Phys 645: 433477. . [6] Pauli, Wolfgang (1940). "The Connection Between Spin and Statistics" (http:/ / web. ihep. su/ dbserv/ compas/ src/ pauli40b/ eng. pdf) (PDF). Phys. Rev 58 (8): 716722. Bibcode1940PhRv...58..716P. doi:10.1103/PhysRev.58.716. . [7] Physics of Atoms and Molecules, B.H. Bransden, C.J.Joachain, Longman, 1983, ISBN 0-582-44401-2 [8] "CODATA Value: electron g factor" (http:/ / physics. nist. gov/ cgi-bin/ cuu/ Value?gem). The NIST Reference on Constants, Units, and Uncertainty. NIST. 2006. . Retrieved 2008-10-18. [9] R.P. Feynman (1985). "Electrons and Their Interactions". QED: The Strange Theory of Light and Matter. Princeton, New Jersey: Princeton University Press. p.115. ISBN0-691-08388-6.

"After some years, it was discovered that this value [g/2] was not exactly 1, but slightly moresomething like 1.00116. This correction was worked out for the first time in 1948 by Schwinger as j*j divided by 2 pi [sic] [where j is the square root of the fine-structure constant], and was due to an alternative way the electron can go from place to place: instead of going directly from one point to another, the electron goes along for a while and suddenly emits a photon; then (horrors!) it absorbs its own photon."
[10] W.J. Marciano, A.I. Sanda (1977). "Exotic decays of the muon and heavy leptons in gauge theories". Physics Letters B67 (3): 303305. Bibcode1977PhLB...67..303M. doi:10.1016/0370-2693(77)90377-X. [11] B.W. Lee, R.E. Shrock (1977). "Natural suppression of symmetry violation in gauge theories: Muon- and electron-lepton-number nonconservation". Physical Review D16 (5): 14441473. Bibcode1977PhRvD..16.1444L. doi:10.1103/PhysRevD.16.1444. [12] K. Fujikawa, R. E. Shrock (1980). "Magnetic Moment of a Massive Neutrino and Neutrino-Spin Rotation". Physical Review Letters 45 (12): 963966. Bibcode1980PhRvL..45..963F. doi:10.1103/PhysRevLett.45.963. [13] N.F. Bell et al. (2005). "How Magnetic is the Dirac Neutrino?". Physical Review Letters 95 (15): 151802. arXiv:hep-ph/0504134. Bibcode2005PhRvL..95o1802B. doi:10.1103/PhysRevLett.95.151802. PMID16241715. [14] Quanta: A handbook of concepts, P.W. Atkins, Oxford University Press, 1974, ISBN 0-19-855493-1 [15] Datta. S and B. Das (1990). "Electronic analog of the electrooptic modulator". Applied Physics Letters 56 (7): 665667. Bibcode1990ApPhL..56..665D. doi:10.1063/1.102730. [16] B. Friedrich, D. Herschbach (2003). "Stern and Gerlach: How a Bad Cigar Helped Reorient Atomic Physics". Physics Today 56 (12): 53. Bibcode2003PhT....56l..53F. doi:10.1063/1.1650229.

References External links


" Spintronics. Feature Article (http://www.sciam.com/article. cfm?articleID=0007A735-759A-1CDD-B4A8809EC588EEDF)" in Scientific American, June 2002. Goudsmit on the discovery of electron spin. (http://www.lorentz.leidenuniv.nl/history/spin/goudsmit.html) Nature: " Milestones in 'spin' since 1896. (http://www.nature.com/milestones/milespin/index.html)" ECE 495N Lecture 36: Spin (http://nanohub.org/resources/6025) Online lecture by S. Datta

Land g-factor

168

Land g-factor
In physics, the Land g-factor is a particular example of a g-factor, namely for an electron with both spin and orbital angular momenta. It is named after Alfred Land, who first described it in 1921. In atomic physics, it is a multiplicative term appearing in the expression for the energy levels of an atom in a weak magnetic field. The quantum states of electrons in atomic orbitals are normally degenerate in energy, with the degenerate states all sharing the same angular momentum. When the atom is placed in a weak magnetic field, however, the degeneracy is lifted. The factor comes about during the calculation of the first-order perturbation in the energy of an atom when a weak uniform magnetic field (that is, weak in comparison to the system's internal magnetic field) is applied to the system. Formally we can write the factor as,[1]

The orbital g-factor is equal to 1, and under the approximation

, the above expression simplifies to

Here, J is the total electronic angular momentum, L is the orbital angular momentum, and S is the spin angular momentum. Because S=1/2 for electrons, one often sees this formula written with 3/4 in place of S(S+1). The quantities gL and gS are other g-factors of an electron. If we wish to know the g-factor for an atom with total atomic angular momentum F=I+J,

This last approximation is justified because mass.

is smaller than

by the ratio of the electron mass to the proton

References
[1] http:/ / hyperphysics. phy-astr. gsu. edu/ HBASE/ quantum/ Lande. html Hyperphysics: Magnetic Interactions and the Land g-Factor

g-factor (physics)

169

g-factor (physics)
For the acceleration-related quantity in mechanics, see g-force. A g-factor (also called g value or dimensionless magnetic moment) is a dimensionless quantity which characterizes the magnetic moment and gyromagnetic ratio of a particle or nucleus. It is essentially a proportionality constant that relates the observed magnetic moment of a particle to the appropriate angular momentum quantum number and the appropriate fundamental quantum unit of magnetism, usually the Bohr magneton or nuclear magneton.

Calculation
Electron g-factors
There are three magnetic moments associated with an electron: One from its spin angular momentum, one from its orbital angular momentum, and one from its total angular momentum (the quantum-mechanical sum of those two components). Corresponding to these three moments are three different g-factors: Electron spin g-factor The most famous of these is the electron spin g-factor (more often called simply the electron g-factor), ge, defined by

where S is the total magnetic moment resulting from the spin of an electron, S is the magnitude of its spin angular momentum, and B is the Bohr magneton. In atomic physics, the electron spin g-factor is often defined as the absolute value or negative of ge: The z-component of the magnetic moment then becomes The value gS is roughly equal to 2.002319, and is known to extraordinary precision.[1][2] The reason it is not precisely two is explained by quantum electrodynamics calculation of the anomalous magnetic dipole moment.[3] Electron orbital g-factor Secondly, the electron orbital g-factor, gL, is defined by

where L is the total magnetic moment resulting from the orbital angular momentum of an electron, L is the magnitude of its orbital angular momentum, and B is the Bohr magneton. The value of gL is exactly equal to one, by a quantum-mechanical argument analogous to the derivation of the classical magnetogyric ratio. For an electron in an orbital with a magnetic quantum number ml, the z-component of the orbital angular momentum is which, since gL = 1, is just Bml

g-factor (physics) Land g-factor Thirdly, the Land g-factor, gJ, is defined by

170

where is the total magnetic moment resulting from both spin and orbital angular momentum of an electron, J = L+S is its total angular momentum, and B is the Bohr magneton. The value of gJ is related to gL and gS by a quantum-mechanical argument; see the article Land g-factor.

Nucleon and nucleus g-factors


Protons, neutrons, and many nuclei have spin and magnetic moments, and therefore associated g-factors. The formula conventionally used is

where is the magnetic moment resulting from the nuclear spin, I is the nuclear spin angular momentum, N is the nuclear magneton, and g is the effective g-factor.

Muon g-factor
The muon, like the electron has a g-factor from its spin, given by the equation

where is the magnetic moment resulting from the muons spin, S is the spin angular momentum, and m is the muon mass.

The fact that the muon g-factor is not quite the same as the electron g-factor is mostly explained by quantum electrodynamics and its calculation of the anomalous magnetic dipole moment. Almost all of the small difference between the two values (99.96% of it) is due to a well-understood lack of a heavy-particle diagrams contributing to the probability for emission of a photon representing the magnetic dipole field, which are present for muons, but not electrons, in QED theory. These are entirely a result of the mass difference between the particles. However, not all of the difference between the g-factors for electrons and muons are exactly explained by the quantum electrodynamics Standard Model. The muon g-factor can, at least in theory, be affected by physics beyond the Standard Model, so it has been measured very precisely, in particular at the Brookhaven National Laboratory. As of November 2006, the experimentally measured value is 2.0023318416(13), compared to the theoretical prediction of 2.0023318361(10).[4] This is a difference of 3.4 standard deviations, suggesting beyond-the-Standard-Model physics may be having an effect.

If supersymmetry is realized in nature, there will be corrections to g-2 of the muon due to loop diagrams involving the new particles. Amongst the leading corrections are those depicted here: a neutralino and a smuon loop , and a chargino and a muon sneutrino loop. This represents an example of "beyond the Standard-Model" physics that might contribute to g-2.

Measured g-factor values

g-factor (physics)

171

Elementary Particle Electron Neutron Proton Muon

g-factor

Uncertainty

2.0023193043622 0.0000000000015 3.82608545 5.585694713 2.0023318414 0.00000090 0.000000046 0.0000000012

|+Currently accepted NIST g-factor values [5] The electron g-factor is one of the most precisely measured values in physics, with its uncertainty beginning at the twelfth decimal place.

Notes and references


[1] Gabrielse, Gerald; Hanneke, David (October 2006). "Precision pins down the electron's magnetism" (http:/ / cerncourier. com/ main/ article/ 46/ 8/ 20). CERN Courier 46 (8): 3537. . [2] Odom, B.; Hanneke, D.; dUrso, B.; Gabrielse, G. (2006). "New measurement of the electron magnetic moment using a one-electron quantum cyclotron". Physical Review Letters 97 (3): 030801. Bibcode2006PhRvL..97c0801O. doi:10.1103/PhysRevLett.97.030801. PMID16907490. [3] Brodsky, S; Franke, V; Hiller, J; McCartor, G; Paston, S; Prokhvatilov, E (2004). "A nonperturbative calculation of the electron's magnetic moment". Nuclear Physics B 703 (12): 333362. arXiv:hep-ph/0406325. Bibcode2004NuPhB.703..333B. doi:10.1016/j.nuclphysb.2004.10.027. [4] Hagiwara, K.; Martin,, A. D.; Nomura, Daisuke; Teubner, T. (2006). "Improved predictions for g-2 of the muon and alpha(QED)(M(Z)**2)". Physics Letters B 649 (23): 173179. arXiv:hep-ph/0611102. Bibcode2007PhLB..649..173H. doi:10.1016/j.physletb.2007.04.012. [5] "CODATA values of the fundamental constants" (http:/ / physics. nist. gov/ cgi-bin/ cuu/ Category?view=html& All+ values. x=80& All+ values. y=11). NIST. .

Zeeman effect
The Zeeman effect ( /zemn/; IPA:[zemn]) is the effect of splitting a spectral line into several components in the presence of a static magnetic field. It is analogous to the Stark effect, the splitting of a spectral line into several components in the presence of an electric field. The Zeeman effect is very important in applications such as nuclear magnetic resonance spectroscopy, electron spin resonance spectroscopy, magnetic resonance imaging (MRI) and Mssbauer spectroscopy. It may also be utilized to improve accuracy in Atomic absorption spectroscopy.

Zeeman splitting of the 5s level of Rb-87, including fine structure and hyperfine structure splitting. Here F = J + I, where I is the nuclear spin. (for Rb-87, I = 3/2)

When the spectral lines are absorption lines, the effect is called inverse Zeeman effect. The Zeeman effect is named after the Dutch physicist Pieter Zeeman.

Zeeman effect

172

Introduction
Without a magnetic field, configurations a, b and c have the same energy, as do d, e and f. The presence of a magnetic field (B) splits the energy levels. Therefore, a line produced by a transition from a, b or c to d, e or f will now be split into several components between different combinations of a, b, c and d, e, f. However, not all transitions will be possible (in the dipole approximation), as governed by the selection rules. Since the distance between the Zeeman sub-levels is proportional to the magnetic field, this effect can be used by astronomers to measure the magnetic field of the Sun and other stars. There is also an anomalous Zeeman effect that appears on transitions where the net spin of the electrons is not 0, the number of Zeeman sub-levels being even instead of odd if there's an uneven number of electrons involved. It was called "anomalous" because the electron spin had not yet been discovered, and so there was no good explanation for it at the time that Zeeman observed the effect. At higher magnetic fields the effect ceases to be linear. At even higher field strength, when the strength of the external field is comparable to the strength of the atom's internal field, electron coupling is disturbed and the spectral lines rearrange. This is called the Paschen-Back effect.

Theoretical presentation
The total Hamiltonian of an atom in a magnetic field is

where

is the unperturbed Hamiltonian of the atom, and

is perturbation due to the magnetic field:

where

is the magnetic moment of the atom. The magnetic moment consists of the electronic and nuclear parts,

however, the latter is many orders of magnitude smaller and will be neglected here. Therefore,

where

is the Bohr magneton,

is the total electronic angular momentum, and

is the Land g-factor. The

operator of the magnetic moment of an electron is a sum of the contributions of the orbital angular momentum and the spin angular momentum where and , with each multiplied by the appropriate gyromagnetic ratio: (the latter is called the anomalous gyromagnetic ratio; the deviation of the

value from 2 is due to Quantum Electrodynamics effects). In the case of the LS coupling, one can sum over all electrons in the atom:

where

and

are the total orbital momentum and spin of the atom, and averaging is done over a state with a is small (less than the fine structure), it can be treated as a perturbation; this is the exceeds the LS coupling significantly (but , in ). In ultrastrong magnetic fields, the magnetic-field interaction may exceed

given value of the total angular momentum. If the interaction term is still small compared to Zeeman effect proper. In the Paschen-Back effect, described below,

which case the atom can no longer exist in its normal meaning, and one talks about Landau levels instead. There are, of course, intermediate cases which are more complex than these limit cases.

Zeeman effect

173

Weak field (Zeeman effect)


If the spin-orbit interaction dominates over the effect of the external magnetic field, conserved, only the total angular momentum then the projection of the spin onto the direction of : thought of as precessing about the (fixed) total angular momentum vector and are not separately is. The spin and orbital angular momentum vectors can be . The (time-)"averaged" spin vector is

and for the (time-)"averaged" orbital vector:

Thus,

Using

and squaring both sides, we get

and: using

and squaring both sides, we get

Combining everything and taking external magnetic field,

, we obtain the magnetic potential energy of the atom in the applied

where the quantity in square brackets is the Land g-factor gJ of the atom ( z-component of the total angular momentum. For a single electron above filled shells the Land g-factor can be simplified into:

and

) and and

is the ,

Example: Lyman alpha transition in hydrogen


The Lyman alpha transition in hydrogen in the presence of the spin-orbit interaction involves the transitions and In the presence of an external magnetic field, the weak-field Zeeman effect splits the 1S1/2 and 2P1/2 states into 2 levels each ( ) and the 2P3/2 state into 4 levels ( ). The Land g-factors for the three levels are: for for for (j=1/2, l=0) (j=1/2, l=1) (j=3/2, l=1).

Note in particular that the size of the energy splitting is different for the different orbitals, because the gJ values are different. On the left, fine structure splitting is depicted. This splitting occurs even in the absence of a magnetic field,

Zeeman effect as it is due to spin-orbit coupling. Depicted on the right is the additional Zeeman splitting, which occurs in the presence of magnetic fields.

174

Strong field (Paschen-Back effect)


The Paschen-Back effect is the splitting of atomic energy levels in the presence of a strong magnetic field. This occurs when an external magnetic field is sufficiently large to disrupt the coupling between orbital ( ) and spin ( ) angular momenta. This effect is the strong-field limit of the Zeeman effect. When , the two effects are equivalent. The effect was named after the German physicists Friedrich Paschen and Ernst E. A. Back. When the magnetic-field perturbation significantly exceeds the spin-orbit interaction, one can safely assume . This allows the expectation values of and to be easily evaluated for a state . The energies are simply:

The above may be read as implying that the LS-coupling is completely broken by the external field. However and are still "good" quantum numbers. Together with the selection rules for an electric dipole transition, i.e., this allows to ignore the spin degree of freedom altogether. As a result, only three spectral lines will be visible, corresponding to the considered. It should be noted that in general (if selection rule. The splitting is independent of the unperturbed energies and electronic configurations of the levels being ), these three components are actually groups of several transitions each, due to the residual spin-orbit coupling. In general, one must now add spin-orbit coupling and relativistic corrections (which are of the same order, known as 'fine structure') as a perturbation to these 'unperturbed' levels. First order perturbation theory with these fine-structure corrections yields the following formula for the Hydrogen atom in the Paschen-Back limit:[1]

Zeeman effect

175

Intermediate field for j = 1/2


In the magnetic dipole approximation, the Hamiltonian which includes both the hyperfine and Zeeman interactions is

To arrive at the Breit-Rabi formula we will include the hyperfine structure (interaction between the electron's spin and the magnetic moment of the nucleus), which is governed by the quantum number , where is the spin angular momentum operator of the nucleus. Alternatively, the derivation could be done with is known as the zero field hyperfine constant and is given in units of Hertz. are the electron and nuclear angular momentum operators. and is the Bohr and can be found via a only. The constant magneton.

classical vector coupling model or a more detailed quantum mechanical calculation to be:

As discussed, in the case of weak magnetic fields, the Zeeman interaction can be treated as a perturbation to the basis. In the high field regime, the magnetic field becomes so large that the Zeeman effect will dominate, and we must use a more complete basis of eigenstates which are superpositions of the or just and since and will be constant , the

within a given level. To get the complete picture, including intermediate field strengths, we must consider basis states. For Hamiltonian can be solved analytically, resulting in the Breit-Rabi formula. Notably, the electric quadrapole interaction is zero for ( ), so this formula is fairly accurate. To solve this system, we note that at all times, the total angular momentum projection between states will change between only will be conserved. Furthermore, since . Therefore, we can define a good basis as:

We now utilize quantum mechanical ladder operators, which are defined for a general angular momentum operator as

These ladder operators have the property

as long as

lies in the range

(otherwise, they return zero). Using ladder operators

and

We can rewrite the Hamiltonian as

Now we can determine the matrix elements of the Hamiltonian:

Solving for the eigenvalues of this matrix, (as can be done by hand, or more easily, with a computer algebra system) we arrive at the energy shifts:

Zeeman effect

176

where

is the splitting (in units of Hz) between two hyperfine sublevels in the absence of magnetic field

is referred to as the 'field strength parameter' (Note: for and should be interpreted as

the square root is an exact square,

). This equation is known as the Breit-Rabi formula and is useful for

systems with one valence electron in an ( ) level.[2][3] Note that index in should be considered not as total angular momentum of the atom but as asymptotic total angular momentum. It is equal to total angular momentum only if (the only exceptions are ). otherwise eigenvectors but equal corresponding different eigenvalues of the Hamiltonian are the superpositions of states with different

Applications
Astrophysics
George Ellery Hale was the first to notice the Zeeman effect in the solar spectra, indicating the existence of strong magnetic fields in sunspots. Such fields can be quite high, on the order of .1 Tesla or higher. Today, the Zeeman effect is used to produce magnetograms showing the variation of magnetic field on the sun.

Magnetic sense in animals


One theory about the magnetic sense of birds assumes that a protein in the retina is changed due to the Zeeman effect.[4]

References
[1] Griffiths, David J. (2004). Introduction to Quantum Mechanics (2nd ed.). Prentice Hall. ISBN0-13-111892-7. OCLC40251748} pg. 247. [2] Woodgate, Elementary Atomic Structure, section 9. [3] first appeared in G. Breit and I. Rabi, Phys. rev. 38, 2082 (1931). [4] The magnetic compass mechanisms of birds and rodents are based on different physical principles (http:/ / rsif. royalsocietypublishing. org/ content/ 3/ 9/ 583. full). Journal of the Royal Society

Zeeman effect on a sunspot spectral line

Historical
Condon, E. U.; G. H. Shortley (1935). The Theory of Atomic Spectra. Cambridge University Press. ISBN0-521-09209-4. (Chapter 16 provides a comprehensive treatment, as of 1935.)
Solar magnetogram

Zeeman, P. (1897). "On the influence of Magnetism on the Nature of the Light emitted by a Substance". Phil. Mag. 43: 226. Zeeman, P. (1897). "Doubles and triplets in the spectrum produced by external magnetic forces". Phil. Mag. 44: 55. Zeeman, P. (11 February 1897). "The Effect of Magnetisation on the Nature of Light Emitted by a Substance" (http://www.nature.com/nature/journal/v55/n1424/abs/055347a0.html). Nature 55 (1424): 347. Bibcode1897Natur..55..347Z. doi:10.1038/055347a0.

Zeeman effect

177

Modern
Feynman, Richard P., Leighton, Robert B., Sands, Matthew (1965). The Feynman Lectures on Physics, Vol. 3. Addison-Wesley. ISBN0-201-02115-3. Forman, Paul (1970). "Alfred Land and the anomalous Zeeman Effect, 1919-1921". Historical Studies in the Physical Sciences 2: 153261. Griffiths, David J. (2004). Introduction to Quantum Mechanics (2nd ed.). Prentice Hall. ISBN0-13-805326-X. Liboff, Richard L. (2002). Introductory Quantum Mechanics. Addison-Wesley. ISBN0-8053-8714-5. Sobelman, Igor I. (2006). Theory of Atomic Spectra. Alpha Science. ISBN1-84265-203-6. Foot, C. J. (2005). Atomic Physics. ISBN0-19-850696-1.

Spin states (d electrons)


Spin states when describing transition metal coordination complexes refers to the potential spin configurations of the metal centers d electrons. In many molecules these spin states vary between high-spin and low-spin configurations. These configurations can be understood through the two major models used to describe coordination complexes; ligand field theory, which is an application of molecular orbital theory to transition metals, and crystal field theory, which has roots in VSEPR theory.[1]

High-spin vs Low-spin
Octahedral complexes
The splitting of the d-orbitals plays an important role in the electron spin state of a coordination complex. There are three factors that affect the : the period of the metal center, the charge of the metal center, and the field strength of the complex's ligands as described by the spectrochemical series. In order for low spin splitting to occur, the energy cost Low-spin [Fe(NO2)6]3 crystal field diagram of placing an electron into an already singly occupied orbital must be less than the cost of placing the additional electron into an eg orbital at an energy cost of . If the energy required to pair two electrons is greater than the energy cost of placing an electron in an eg, , high spin splitting occurs. If the separation between the orbitals is large, then the lower energy orbitals are completely filled before population of the higher orbitals according to the Aufbau principle. Complexes such as this are called "low-spin" since filling an orbital matches electrons and reduces the total electron spin. If the separation between the orbitals is small enough then it is easier to put electrons into the higher energy orbitals than it is to put two into the same low-energy orbital, because of the repulsion resulting from matching two electrons in the same orbital. So, one electron is put into each of the five d-orbitals before any pairing occurs in accord with Hund's rule resulting in what is known as a "high-spin" complex. Complexes such as this are called "high-spin" since populating the upper orbital avoids matches between electrons with opposite spin.

Spin states (d electrons)

178

Within a transition metal group moving down the series corresponds with an increase in . The observed result is larger splitting for complexes in octahedral geometries based around transition metal centers of the second or third row, periods 5 and 6 respectively. This splitting is generally large enough that these complexes do not exist as high-spin state. This is true High-spin [FeBr6]3 crystal field diagram even when the metal center is coordinated to weak field ligands. It is only octahedral coordination complexes which are centered around first row transition metals that fluctuate between high and low-spin states. The charge of the metal center plays a role in the ligand field and the splitting. For example, Fe2+ and Co3+ are both d6; however, the higher charge of Co3+ creates a stronger ligand field than Fe2+. All other things being equal, Fe2+ is more likely to be high spin than Co3+. Ligands also affect the magnitude of splitting of the d-orbitals according to their field strength as described by the spectrochemical series. Strong-field ligands, such as CN and CO, increase the splitting and are more likely to be low-spin. Weak-field ligands, such as I and Br cause a smaller splitting and are more likely to be high-spin.

Tetrahedral complexes
The splitting energy for tetrahedral metal complexes (four ligands), tet is smaller than that for an octahedral complex. Therefore, it is rare to have a tet large enough to cause electrons to pair before filling high orbitals. Thus, tetrahedral complexes are usually high-spin. "There are no known ligands powerful enough to produce the strong-field case in a tetrahedral complex" (Transition metals and Coordination Chemistry:The Crystal field Model by Steven S. Zumdahl. Chemical Principles)

Square planar complexes


Most spin state transitions are between the same geometry, namely octahedral. However, in the case of d8 complexes is a shift in geometry between spin states. There is no possible difference between the high and low-spin states in the d8 octahedral complexes, however d8 complexes are able to shift from paramagnetic tetrahedral geometry to a diamagnetic low-spin square planar geometry.

Ligand field theory vs Crystal field theory


The rationale for why the spin states exist according to ligand field theory is essentially the same as the crystal field theory explanation. However the explanation of why the orbitals split is different accordingly with each model and requires translation.

High-spin and low-spin systems


The first d electron count (special version of electron configuration) with the possibility of holding a high spin or low spin state is octahedral d4 since it has more than the 3 electrons to fill the non bonding d orbitals according to ligand field theory or the stabilized d orbitals according to crystal field splitting. The spin state of the complex also affects an atom's ionic radius.[2] d4 Octahedral high-spin: 4 unpaired electrons, paramagnetic, substitutionally labile. Includes Cr2+ ionic radius 80 pm, Mn3+ ionic radius 64.5 pm.

Spin states (d electrons) Octahedral low-spin: 2 unpaired electrons, paramagnetic, substitutionally inert. Includes Cr2+ ionic radius 73 pm, Mn3+ ionic radius 58 pm. d5 Octahedral high-spin: 5 unpaired electrons, paramagnetic, substitutionally labile. Includes Fe3+ ionic radius 64.5 pm. Octahedral low-spin: 1 unpaired electron, paramagnetic, substitutionally inert. Includes Fe3+ ionic radius 55 pm. d6 Octahedral high-spin: 4 unpaired electrons, paramagnetic, substitutionally labile. Includes Fe2+ ionic radius 78 pm, Co3+ ionic radius 61 pm. Octahedral low-spin: no unpaired electrons, diamagnetic, substitutionally inert. Includes Fe2+ ionic radius 62 pm, Co3+ ionic radius 54.5 pm, Ni4+ ionic radius 48 pm. d7 Octahedral high-spin: 3 unpaired electrons, paramagnetic, substitutionally labile. Includes Co2+ ionic radius 74.5 pm, Ni3+ ionic radius 60 pm. Octahedral low-spin:1 unpaired electron, paramagnetic, substitutionally labile. Includes Co2+ ionic radius 65 pm, Ni3+ ionic radius 56 pm. d8 Octahedral high-spin: 2 unpaired electrons, paramagnetic, substitutionally labile. Includes Ni2+ ionic radius 69 pm. Square planar low-spin: no unpaired electrons, diamagnetic, substitutionally inert. Includes Ni2+ ionic radius 49 pm.

179

References
[1] Miessler, Gary L.; Donald A. Tarr (1998). Inorganic Chemistry (2nd edition). Upper Saddle River, New Jersey: Pearson Education, Inc. Pearson Prentice Hall. ISBN0138418918. [2] Revised effective ionic radii and systematic studies of interatomic distances in halides and chalcogenides Shannon R.D. Acta Cryst. A32 751-767 (1976) doi:10.1107/S0567739476001551

Spinorbit interaction

180

Spinorbit interaction
In quantum physics, the spinorbit interaction (also called spinorbit effect or spinorbit coupling) is any interaction of a particle's spin with its motion. The first and best known example of this is that spin-orbit interaction causes shifts in an electron's atomic energy levels due to electromagnetic interaction between the electron's spin and the magnetic field generated by the electron's orbit around the nucleus. This is detectable as a splitting of spectral lines. A similar effect, due to the relationship between angular momentum and the strong nuclear force, occurs for protons and neutrons moving inside the nucleus, leading to a shift in their energy levels in the nucleus shell model. In the field of spintronics, spin-orbit effects for electrons in semiconductors and other materials are explored and put to useful work.

Spinorbit interaction in atomic energy levels


Using some semiclassical electrodynamics and non-relativistic quantum mechanics, in this section we present a relatively simple and quantitative description of the spin-orbit interaction for an electron bound to an atom, up to first order in perturbation theory. This gives results that agree reasonably well with observations. A more rigorous derivation of the same result would start with the Dirac equation, and achieving a more precise result would involve calculating small corrections from quantum electrodynamics.

Energy of a magnetic moment


The energy of a magnetic moment in a magnetic field is given by:

where is the magnetic moment of the particle and B is the magnetic field it experiences.

Magnetic field
We shall deal with the magnetic field first. Although in the rest frame of the nucleus, there is no magnetic field, there is one in the rest frame of the electron. Ignoring for now that this frame is not inertial, in SI units we end up with the equation

where v is the velocity of the electron and E the electric field it travels through. Now we know that E is radial so we can rewrite . Also we know that the momentum of the electron . Substituting this in and

changing the order of the cross product gives:

Next, we express the electric field as the gradient of the electric potential

. Here we make the central

field approximation, that is, that the electrostatic potential is spherically symmetric, so is only a function of radius. This approximation is exact for hydrogen, and indeed hydrogen-like systems. Now we can say

where we get

is the potential energy of the electron in the central field, and e is the elementary charge. Now we . Putting it all together

remember from classical mechanics that the angular momentum of a particle

Spinorbit interaction

181

It is important to note at this point that B is a positive number multiplied by L, meaning that the magnetic field is parallel to the orbital angular momentum of the particle.

Magnetic Moment of the Electron


The magnetic moment of the electron is

where Here,

is the spin angular momentum vector,

is the Bohr magneton and

is the electron spin g-factor.

is a negative constant multiplied by the spin, so the magnetic moment is antiparallel to the spin angular

momentum. The spin-orbit potential consists of two parts. The Larmor part is connected to interaction of the magnetic moment of electron with magnetic field of nucleus in the co-moving frame of electron. The second contribution is related to Thomas precession.

Larmor interaction energy


The Larmor interaction energy is

Substituting in this equation expressions for the magnetic moment and the magnetic field, one gets

Now, we have to take into account Thomas precession correction for the electron's curved trajectory.

Thomas interaction energy


In 1926 Llewellyn Thomas relativistically recomputed the doublet separation in the fine structure of the atom.[1]. Thomas precession rate, , is related to the angular frequency of the orbital motion, , of a spinning particle as follows [2][3]

where

is the Lorentz factor of moving particle. The Hamiltonian producing the spin precession

is given by

To the first order in

, we obtain

Total interaction energy


The total spin-orbit potential in an external electrostatic potential takes the form

The net effect of Thomas precession is the reduction of the Larmor interaction energy by factor 1/2 which came to be known as the Thomas half.

Spinorbit interaction

182

Evaluating the energy shift


Thanks to all the above approximations, we can now evaluate the detailed energy shift in this model. In particular, we wish to find a basis that diagonalizes both H0 (the non-perturbed Hamiltonian) and H. To find out what basis this is, we first define the total angular momentum operator

Taking the dot product of this with itself, we get

(since L and S commute), and therefore

It can be shown that the five operators H0, J, L, S, and Jz all commute with each other and with H. Therefore, the basis we were looking for is the simultaneous eigenbasis of these five operators (i.e., the basis where all five are diagonal). Elements of this basis have the five quantum numbers: n (the "principal quantum number") j (the "total angular momentum quantum number"), l (the "orbital angular momentum quantum number"), s (the "spin quantum number"), and jz (the "z-component of total angular momentum"). To evaluate the energies, we note that

for hydrogenic wavefunctions (here

is the Bohr radius divided by the nuclear charge Z); and

Final Energy Shift


We can now say

where

For hydrogen, we can write the explicit result

For any hydrogen-like atom with Z protons

Spinorbit interaction

183

References
[1] L. H. Thomas, The motion of the spinning electron, Nature (London), 117, 514 (1926). [2] L. Fppl and P. J. Daniell, Zur Kinematik des Born'schen starren Krpers, Nachrichten von der Kniglichen Gesellschaft der Wissenschaften zu Gttingen, 519 (1913). [3] C. Mller, The Theory of Relativity, (Oxford at the Claredon Press, London, 1952).

Textbooks
E. U. Condon and G. H. Shortley (1935). The Theory of Atomic Spectra. Cambridge University Press. ISBN0-521-09209-4. D. J. Griffiths (2004). Introduction to Quantum Mechanics (2nd edition). Prentice Hall. Landau, Lev; L. M. Lifshitz. " Theory, Volume 3. 72. Fine structure of atomic levels". Quantum Mechanics: Non-Relativistic

Azimuthal quantum number


The azimuthal quantum number is a quantum number for an atomic orbital that determines its orbital angular momentum and describes the shape of the orbital. The azimuthal quantum number is the second of a set of quantum numbers which describe the unique quantum state of an electron (the others being the principal quantum number, following spectroscopic notation, the magnetic quantum number, and the spin quantum number). It is also known as the orbital angular momentum quantum number or second quantum number, and is symbolized as (lower-case L).

Derivation
Associated with the energy states of the electrons of an atom is a set of four quantum numbers: n, , m, and ms. These specify the complete and unique quantum state of a single electron in an atom, and make up its wavefunction or orbital. The wavefunction of the Schrdinger wave equation reduces to three equations that when solved, lead to the first three quantum numbers. Therefore, the equations for the first three quantum numbers are all interrelated. The azimuthal quantum number arose in the solution of the polar part of the wave equation as shown below. To aid understanding of this concept of the azimuth, it may also prove helpful to review spherical coordinate systems, and/or other alternative mathematical coordinate systems besides the cartesian coordinate system. Generally, the spherical coordinate system works best with spherical models, the cylindrical system with cylinders, the cartesian with general volumes, etc. An atomic electron's angular momentum, L, is related to its quantum number by the following equation: where is the reduced Planck's constant, L2 is the orbital angular momentum operator and

is the wavefunction of

the electron. The quantum number is always a nonnegative integer: 0,1,2,3, etc. (see angular momentum quantization). While many introductory textbooks on quantum mechanics will refer to L by itself, L has no real meaning except in its use as the angular momentum operator. When referring to angular momentum, it is best to simply use the quantum number . The energy of any wave is the frequency multiplied by Planck's constant. This causes the wave to display particle-like packets of energy called quanta. To show each of the quantum numbers in the quantum state, the formulae for each quantum number include Planck's reduced constant which only allows particular or discrete or quantized energy levels. This behavior manifests itself as the "shape" of the orbital.

Azimuthal quantum number

184

Atomic orbitals have distinctive shapes denoted by letters. In the illustration, the letters s, p, and d describe the shape of the atomic orbital. Their wavefunctions take the form of spherical harmonics, and so are described by Legendre polynomials. The various orbitals relating to different values of are sometimes called sub-shells, and (mainly for historical reasons) are referred to by letters, as follows:

The atomic orbital wavefunctions of a hydrogen atom. The principal quantum number is at the right of each row and the azimuthal quantum number is denoted by letter at top of each column.

Letter Max electrons 0 s 1 p 2 d 3 f 4 g 5 h 6 i 2 6 10 14 18 22 26 sphere two dumbbells

Shape

Name sharp principal diffuse

four dumbbells or unique shape one

eight dumbbells or unique shape two fundamental

A mnemonic for the order of the "sub-shells" is some poor dumb fool. Another mnemonic for the order of the "sub-shells" is silly professors dance funny. The letters after the f sub-shell just follow f in alphabetical order. Each of the different angular momentum states can take 2(2 + 1) electrons. This is because the third quantum number m (which can be thought of loosely as the quantized projection of the angular momentum vector on the z-axis) runs from to in integer units, and so there are 2 + 1 possible states. Each distinct n,,m orbital can be occupied by two electrons with opposing spins (given by the quantum number ms), giving 2(2 + 1) electrons overall. Orbitals with higher than given in the table are perfectly permissible, but these values cover all atoms so far discovered. For a given value of the principal quantum number n, the possible values of range from 0 to n 1; therefore, the n = 1 shell only possesses an s subshell and can only take 2 electrons, the n = 2 shell possesses an s and a p subshell and can take 8 electrons overall, the n = 3 shell possesses s, p and d subshells and has a maximum of 18 electrons, and so on. Generally speaking, the maximum number of electrons in the nth energy level is 2n2. The angular momentum quantum number, , governs the number of planar nodes going through the nucleus. A planar node can be described in an electromagnetic wave as the midpoint between crest and trough, which has zero magnitude. In an s orbital, no nodes go through the nucleus, therefore the corresponding azimuthal quantum number takes the value of 0. In a p orbital, one node traverses the nucleus and therefore has the value of 1. L has the value . Depending on the value of n, there is an angular momentum quantum number and the following series. The wavelengths listed are for a hydrogen atom:

Azimuthal quantum number n = 1, L = 0, Lyman series (ultraviolet) n = 2, L = 2, Balmer series (visible) n = 3, L = 6, Ritz-Paschen series (short wave infrared) n = 5, L = 25, Pfund series (long wave infrared).

185

Addition of quantized angular momenta


Given a quantized total angular momentum and , associated with its magnitude can range from to in integer steps where which is the sum of two individual quantized angular momenta

the quantum number and

are quantum numbers corresponding to the magnitudes of the individual angular momenta.

Total angular momentum of an electron in the atom


Due to the spin-orbit interaction in the atom, the orbital angular momentum no longer commutes with the Hamiltonian, nor does the spin. These therefore change over time. However the total angular momentum J does commute with the Hamiltonian and so is constant. J is defined through

L being the orbital angular momentum and S the spin. The total angular momentum satisfies the same commutation relations as orbital angular momentum, namely

from which follows

where Ji stand for Jx, Jy, and Jz. The quantum numbers describing the system, which are constant over time, are now j and mj, defined through the action of J on the wavefunction

So that j is related to the norm of the total angular momentum and mj to its projection along a specified axis. As with any angular momentum in quantum mechanics, the projection of J along other axes cannot be co-defined with Jz, because they do not commute. Relation between new and old quantum numbers j and mj, together with the parity of the quantum state, replace the three quantum numbers , m and ms (the projection of the spin along the specified axis). The former quantum numbers can be related to the latter. Furthermore, the eigenvectors of j, mj and parity, which are also eigenvectors of the Hamiltonian, are linear combinations of the eigenvectors of , m and ms.

List of angular momentum quantum numbers


Intrinsic (or spin) angular momentum quantum number, or simply spin quantum number orbital angular momentum quantum number (the subject of this article) magnetic quantum number, related to the orbital momentum quantum number total angular momentum quantum number

Azimuthal quantum number

186

History
The azimuthal quantum number was carried over from the Bohr model of the atom, and was posited by Arnold Sommerfeld[1]. The Bohr model was derived from spectroscopic analysis of the atom in combination with the Rutherford atomic model. The lowest quantum level was found to have an angular momentum of zero. To simplify the mathematics, orbits were considered as oscillating charges in one dimension and so described as "pendulum" orbits. In three-dimensions the orbit becomes spherical without any nodes crossing the nucleus, similar to a skipping rope that oscillates in one large circle.

References
[1] Eisberg, Robert (1974). Quantum Physics of Atoms, Molecules, Solids, Nuclei and Particles. New York: John Wiley & Sons Inc. pp.114117. ISBN978-0471234647.

External links
Development of the Bohr atom (http://galileo.phys.virginia.edu/classes/252/Bohr_Atom/Bohr_Atom.html) NOTE ON "PENDULUM" ORBITS IN ATOMIC MODELS (http://www.pubmedcentral.gov/picrender. fcgi?tool=pmcentrez&blobtype=pdf&artid=1085028) Pictures of atomic orbitals (http://itl.chem.ufl.edu/ao_pict/ao_pict.html) Detailed explanation of the Orbital Quantum Number l (http://www.src.wits.ac.za/pages/teaching/Connell/ phys284/2005/lecture-03/lecture_03/node7.html) The azimuthal equation explained (http://hyperphysics.phy-astr.gsu.edu/hbase/quantum/hydazi.html#c1)

Principal quantum number


In the principal quantum number symbolized as n is the first of a set of quantum numbers (which includes: the principal quantum number, the azimuthal quantum number, the magnetic quantum number, and the spin quantum number) of an atomic orbital. The principal quantum number can only have positive integer values. As n increases, the orbital becomes larger and the electron spends more time farther from the nucleus. As n increases, the electron is also at a higher potential energy and is therefore less tightly bound to the nucleus. This is the only quantum number introduced by the Bohr model. For an analogy, one could imagine a multistoried building with an elevator structure. The building has an integer number of floors, and a (well-functioning) elevator which can only stop at a particular floor. Furthermore the elevator can only travel an integer number of levels. As with the principal quantum number, higher numbers are associated with higher potential energy. Of course beyond this point the analogy breaks down. In the case of elevators the potential energy is gravitational but with the quantum number it is electromagnetic. The gains and losses in energy are approximate with the elevator, but precise with quantum state. The elevator ride from floor to floor is continuous whereas quantum transitions are discontinuous. Finally the constraints of elevator design are imposed by the requirements of architecture, but quantum behavior reflects fundamental laws of physics.

Principal quantum number

187

Derivation
There are a set of quantum numbers associated with the energy states of the atom. The four quantum numbers n, , m, and s specify the complete and unique quantum state of a single electron in an atom called its wavefunction or orbital. Two electrons belonging to the same atom can not have the same four quantum numbers, due to the Pauli exclusion principle. The wavefunction of the Schrdinger wave equation reduces to the three equations that when solved lead to the first three quantum numbers. Therefore, the equations for the first three quantum numbers are all interrelated. The principle quantum number arose in the solution of the radial part of the wave equation as shown below. The Schrdinger wave equation describes energy eigenstates having corresponding real numbers En with a definite total energy which the value of En defines. The bound state energies of the electron in the hydrogen atom are given by:

The parameter n can take only positive integer values. The concept of energy levels and notation was utilized from the earlier Bohr model of the atom. Schrdinger's equation developed the idea from a flat two-dimensional Bohr atom to the three-dimensional wave function model. In the Bohr model, the allowed orbits were derived from quantized (discrete) values of orbital angular momentum, L according to the equation

where n = 1,2,3, and is called the principle quantum number, and h is Planck's constant. This formula is not correct in quantum mechanics as the angular momentum magnitude is described by the azimuthal quantum number, but the energy levels are accurate and classically they correspond to the sum of potential and kinetic energy of the electron. The principle quantum number n represents the relative overall energy of each orbital, and the energy of each orbital increases as the distance from the nucleus increases. The sets of orbitals with the same n value are often referred to as electron shells or energy levels. The minimum energy exchanged during any wave-matter interaction is the wave frequency multiplied by Planck's constant. This causes the wave to display particle-like packets of energy called quanta. The difference between energy levels that have different n determine the Emission spectrum of the element. In the notation of the periodic table, the main shells of electrons are labeled: K (n = 1), L (n = 2), M (n = 3), etc. based on the principle quantum number. The principle quantum number is related to the radial quantum number, nr, by: where is the azimuthal quantum number and nr is equal to the number of nodes in the radial wavefunction.

Principal quantum number

188

External references
Periodic Table Applet: showing principal and azimuthal quantum number for each element [1]

References
[1] http:/ / www. colorado. edu/ physics/ 2000/ applets/ a2. html

Spin quantum number


In atomic physics, the spin quantum number is a quantum number that parameterizes the intrinsic angular momentum (or spin angular momentum, or simply spin) of a given particle. The spin quantum number is the fourth of a set of quantum numbers (the principal quantum number, the azimuthal quantum number, the magnetic quantum number, and the spin quantum number), which describe the unique quantum state of an electron and is designated by the letters.

Derivation
As a quantized angular momentum, (see angular momentum quantum number) it holds that

where is the quantized spin vector is the norm of the spin vector is the spin quantum number associated with the spin angular momentum is the reduced Planck constant. Given an arbitrary directionz (usually determined by an external magnetic field) the spin z-projection is given by

where ms is the secondary spin quantum number, ranging from s to +s in steps of one. This generates 2 s + 1 different values of ms. The allowed values for s are non-negative integers or half-integers. Fermions (such as the electron, proton or neutron) have half-integer values, whereas bosons (e.g., photon, mesons) have integer spin values.

Algebra
The algebraic theory of spin is a carbon copy of the Angular momentum in quantum mechanics theory. First of all, spin satisfies the fundamental commutation relation: , where lmn is the (antisymmetric) Levi-Civita symbol. This means that it is impossible to know two coordinates of the spin at the same time because of the restriction of the Uncertainty principle. Next, the eigenvectors of and satisfy:

Spin quantum number where operators. are the creation and annihilation (or "raising" and "lowering" or "up" and "down")

189

Electron spin
Early attempts to explain the behavior of electrons in atoms focused on solving the Schrdinger wave equation for the hydrogen atom, the simplest possible case, with a single electron bound to the atomic nucleus. This was successful in explaining many features of atomic spectra. The solutions required each possible state of the electron to be described by three "quantum numbers". These were identified as, respectively, the electron "shell" number n, the "orbital" number l, and the "orbital angular momentum" number m. Angular momentum is a so-called "classical" concept measuring the momentum of a mass in circular motion about a point. The shell numbers start at 1 and increase indefinitely. Each shell of number n contains n orbitals. Each orbital is characterized by its number l, where l takes integer values from 0 to n1, and its angular momentum number m, where m takes integer values from +l to l. By means of a variety of approximations and extensions, physicists were able to extend their work on hydrogen to more complex atoms containing many electrons. Atomic spectra measure radiation absorbed or emitted by electrons "jumping" from one "state" to another, where a state is represented by values of n, l, and m. The so-called "Transition rule" limits what "jumps" are possible. In general, a jump or "transition" is allowed only if all three numbers change in the process. This is because a transition will be able to cause the emission or absorption of electromagnetic radiation only if it involves a change in the electromagnetic dipole of the atom. However, it was recognized in the early years of quantum mechanics that atomic spectra measured in an external magnetic field (see Zeeman effect) cannot be predicted with just n, l, and m. A solution to this problem was suggested in early 1925 by George Uhlenbeck and Samuel Goudsmit, students of Paul Ehrenfest (who rejected the idea), and independently by Ralph Kronig, one of Land's assistants. Uhlenbeck, Goudsmit, and Kronig introduced the idea of the self-rotation of the electron, which would naturally give rise to an angular momentum vector in addition to the one associated with orbital motion (quantum numbers l and m). The spin angular momentum is characterized by a quantum number; s = 1/2 specifically for electrons. In a way analogous to other quantized angular momenta, L, it is possible to obtain an expression for the total spin angular momentum:

where is the reduced Planck constant. The hydrogen spectra fine structure is observed as a doublet corresponding to two possibilities for the z-component of the angular momentum, where for any given directionz:

whose solution has only two possible z-components for the electron. In the electron, the two different spin orientations are sometimes called "spin-up" or "spin-down". The spin property of an electron would give rise to magnetic moment, which was a requisite for the fourth quantum number. The electron spin magnetic moment is given by the formula:

where

Spin quantum number e is the charge of the electron g is the Land g-factor and by the equation:

190

where

is the Bohr magneton.

When atoms have even numbers of electrons the spin of each electron in each orbital has opposing orientation to that of its immediate neighbor(s). However, many atoms have an odd number of electrons or an arrangement of electrons in which there is an unequal number of "spin-up" and "spin-down" orientations. These atoms or electrons are said to have unpaired spins that are detected in electron spin resonance.

Detection of spin
When lines of the hydrogen spectrum are examined at very high resolution, they are found to be closely spaced doublets. This splitting is called fine structure, and was one of the first experimental evidences for electron spin. The direct observation of the electron's intrinsic angular momentum was achieved in the SternGerlach experiment.

The SternGerlach experiment


The theory of spatial quantization of the spin moment of the momentum of electrons of atoms situated in the magnetic field needed to be proved experimentally. In 1920 (two years before the theoretical description of the spin was created) Otto Stern and Walter Gerlach observed it in the experiment they conducted. The atoms of silver from the source that was the furnace with boiling silver were leaded to the vacuum space. There (thanks to the thin slides) the flat beam of those atoms was created. Then the beam got into non-homogeneous magnetic field and incidenced a photographic plate. Using classic physical laws we would expect the single picture of the beam on the plate. Whereas the beam of the atoms passing through not homogeneous magnetic field undergoes splitting. That is why Otto Stern and Walter Gerlach received the two lines on the photographic plate. The phenomena can be explained with the spatial quantization of the spin moment of momentum. In atoms the electrons are typically located in such way that in each pair of electrons there is one of the upward spin and one of the downward spin. So the whole spin of such pair is equal to zero. But, in the atom of silver on the outer shell, there is a single electron whose spin is not balanced by any electron. The circulation causes some magnetic dipole moment (it's like it was a very small magnet). There is a force moment in the magnetic field influencing the dipole that is turning it until its position is the same as the direction of the field B. There is some other force influencing the dipole in the field. When the dipole is directed the same as the magnetic field then the dipole is pulled by that force toward the strongest field. But, if the dipole is directed opposite to the field's, it is pushed away from the strongest field. So the atom of silver having one electron on the outer shell can be pulled in or out the space of a strongest magnetic field, what depends on the value of the magnetic spin quantum number. When the spin of the electron is equal +1/2 the atom is pulled out and when the spin is equal 1/2 the atom is pulled in. So during passing through the non-homogeneous magnetic field the beam of the atoms of silver undergoes splitting into the two beams. Each of them consists of atoms whose outer electrons are of the same spin. In 1927 Phipps and Taylor conducted a similar experiment. This time they used atoms of hydrogen, not silver. They also observed that the beam of atoms undergoes splitting into two beams. Later scientists conducted experiments using other atoms that have only one electron on the outer shell (copper, gold, sodium, potassium). Every time there were two lines achieved on the photographic plate. In the atom, not only electrons have spin: The atomic nucleus also may have spin. But protons and neutrons are much heavier than electrons (about 1836 times), and the magnetic dipole moment is inversely proportional to the

Spin quantum number mass. So the nuclear magnetic dipole momentum is much smaller than that of the whole atom. This small magnetic dipole was later measured by Stern, Frisch and Easterman.

191

Dirac equation solves spin


When the idea of electron spin was first introduced in 1925, even Wolfgang Pauli had trouble accepting Ralph Kronig's model. The problem was not that a rotating charged particle would have given rise to a magnetic field but that the electron was so small that the equatorial speed of the electron would have to be greater than the speed of light for the magnetic moment to be of the observed strength. In 1930, Paul Dirac developed a new version of the Wave Equation which was relativistically invariant (unlike Schrdinger's one), and predicted the magnetic moment correctly, and at the same time treated the electron as a point particle. In the Dirac equation all four quantum numbers including the additional quantum number, s arose naturally during its solution.

External references
Full treatment of Spin--including origins, evolution of Spin Theory, and details of the Spin equations [1]

References
[1] http:/ / math. ucr. edu/ home/ baez/ spin/ spin. html

Total angular momentum quantum number


Further information: Azimuthal quantum number#Addition of quantized angular momenta In quantum mechanics, the total angular quantum momentum number parameterises the total angular momentum of a given particle, by combining its orbital angular momentum and its intrinsic angular momentum (i.e., its spin). If s is the particle's spin angular momentum and its orbital angular momentum vector, the total angular momentum j is

The associated quantum number is the main total angular momentum quantum number j. It can take the following range of values, jumping only in integer steps:

where is the azimuthal quantum number (parameterizing the orbital angular momentum) and s is the spin quantum number (parameterizing the spin). The relation between the total angular momentum vector j and the total angular momentum quantum number j is given by the usual relation (see angular momentum quantum number)

the vector's z-projection is given by

where mj is the secondary total angular momentum quantum number. It ranges from j to +j in steps of one. This generates 2j + 1 different values of mj. The total angular momentum corresponds to the Casimir invariant of the Lie algebra so(3) of the three-dimensional rotation group.

Total angular momentum quantum number

192

References
Griffiths, David J. (2004). Introduction to Quantum Mechanics (2nd ed.). Prentice Hall. ISBN0-13-805326-X.

External links
Vector model of angular momentum [1] LS and jj coupling [2]

References
[1] http:/ / hyperphysics. phy-astr. gsu. edu/ hbase/ quantum/ vecmod. html [2] http:/ / hyperphysics. phy-astr. gsu. edu/ hbase/ atomic/ lcoup. html#c1

Angular momentum operator


In quantum mechanics, the angular momentum operator is one of several related operators analogous to classical angular momentum. The angular momentum operator plays a central role in the theory of atomic physics and other quantum problems involving rotational symmetry. In both classical and quantum mechanical systems, angular momentum (together with linear momentum and energy) is one of the three fundamental properties of motion.[1] There are several angular momentum operators: total angular momentum (usually denoted J), orbital angular momentum (usually denoted L), and spin angular momentum (spin for short, usually denoted S). The term "angular momentum operator" can (confusingly) refer to either the total or the orbital angular momentum. Total angular momentum is always conserved, due to Noether's theorem.

Spin, orbital, and total angular momentum


The classical definition of angular momentum is . This can be carried over to quantum mechanics, by reinterpreting r as the quantum position operator and p as the quantum momentum operator. L is then an operator, specifically called the orbital angular momentum operator. Specifically, L is a vector operator, meaning , where Lx, Ly, Lz are three different operators. However, there is another type of angular momentum, called spin angular momentum (more often shortened to spin), represented by the spin operator S. Almost all elementary particles have spin. Spin is often depicted as a particle literally spinning around an axis, but this is a misleading and inaccurate picture: Spin is an intrinsic property of a particle, unrelated to any sort of motion in space. All elementary particles have a characteristic spin, for example electrons always have "spin 1/2" while photons always have "spin 1". Finally, there is total angular momentum J, which combines both the spin and orbital angular momentum of a particle or system:

Conservation of angular momentum states that J for a closed system, or J for the whole universe, is conserved. However, L and S are not generally conserved. For example, the spinorbit interaction allows angular momentum to transfer back and forth between L and S, with the total J remaining constant.

Angular momentum operator

193

Orbital angular momentum operator


Orbital angular momentum L is mathematically defined as the cross product of a wave function's position operator (r) and momentum operator (p):

This is analogous to the definition of angular momentum in classical physics. In the special case of a single particle with no electric charge and no spin, the angular momentum operator can be written in the position basis as a single vector equation:

where is the gradient operator.

Commutation relations
Commutation relations between components
The orbital angular momentum operator is a vector operator, meaning it can be written in terms of its vector components . The components have the following commutation relations with each other:[2]

or in symbols, , where lmn denotes the Levi-Civita symbol, and l,m,n are Cartesian coordinates (each can be x, y or z), and [ , ] is the commutator . These can be proved as a direct consequence of the canonical commutation relations is the Kronecker delta. There is an analogous relationship in classical physics: , where lm

where

is the Poisson bracket.

The same commutation relations apply for the other angular momentum operators (spin and total angular momentum):[3] . These can be assumed to hold in analogy with L. Alternatively, they can be derived as discussed below. These commutation relations mean that L has the mathematical structure of a Lie algebra. In this case, the Lie algebra is SU(2) or SO(3), the rotation group in three dimensions. The same is true of J and S. The reason is discussed below. These commutation relations are relevant for measurement and uncertainty, as discussed further below.

Angular momentum operator

194

Commutation relations involving vector magnitude


Like any vector, a magnitude can be defined for the orbital angular momentum operator: . L is another quantum operator. It commutes with the components of L:
2

This can be derived starting from the Mathematically,

commutation relations in the previous section.

is a Casimir invariant of the Lie algebra spanned by L.

The same commutation relations apply for the other angular momentum operators (spin and total angular momentum): .

Uncertainty principle
In general, in quantum mechanics, when two observable operators do not commute, they are called incompatible observables. Two incompatible observables cannot be measured simultaneously; instead they satisfy an uncertainty principle. The more accurately one observable is known, the less accurately the other one can be known. Just as there is an uncertainty principle relating position and momentum, there are uncertainty principles for angular momentum. The RobertsonSchrdinger relation gives the following uncertainty principle:

where

is the standard deviation in the measured values of X and

denotes the expectation value of X. This

inequality is also is true if x,y,z are rearranged, or if L is replaced by J or S. Therefore, two orthogonal components of angular momentum cannot be simultaneously known or measured, except in special cases such as . It is, however, possible to simultaneously measure or specify L2 and any one component of L; for example, L2 and Lz. This is often useful, and the values are characterized by azimuthal quantum number and magnetic quantum number, as discussed further below.

Quantization
In quantum mechanics, angular momentum is quantized that is, it cannot vary continuously, but only in "quantum leaps" between certain allowed values. For any system, the following restrictions on measurement results apply, where is reduced Planck constant:
If you measure... Lz , where The result can be... Notes

m is sometimes called "magnetic quantum number". This same quantization rule holds for any component of L, e.g. Lx or Ly. [4] This rule is sometimes called spatial quantization. For Sz, m is sometimes called "spin projection quantum number". For Jz, m is sometimes called "total angular momentum projection quantum number". This same quantization rule holds for any component of S or J, e.g. Sx or Jy.

Sz or Jz

, where

Angular momentum operator

195

, where

L2 is defined by

is sometimes called "azimuthal quantum number" or "orbital quantum number". , where s is called spin quantum number or just "spin". For example, a spin- particle is a particle where s=. j is sometimes called "total angular momentum quantum number". for and (See above for terminology.) (See above for terminology.)

, where and simultaneously where for , and

and simultaneously where

for

, and

for and

and simultaneously where

for

, and

for and

(See above for terminology.)

Derivation using ladder operators


A common way to derive the quantization rules above is the method of ladder operators.[5] The ladder operators are defined:

Suppose a state and definite value of prove that or decreased by

is a state in the simultaneous eigenbasis of and a single, ). Then using the commutation relations, one can and are also in the simultaneous , but where is increased
In this standing wave on a circular string, the circle is broken into exactly 8 wavelengths. A standing wave like this can have 0,1,2, or any integer number of wavelengths around the circle, but it cannot have a non-integer number of wavelengths like 8.3. In quantum mechanics, angular momentum is quantized for a similar reason.

(i.e., a state with a single, definite value of

eigenbasis, with the same value of

, respectively. (It is also possible that one or both of

these vectors is the zero vector.) Next, consider the sequence ("ladder") of states

Each nonzero state has a value of

which is

greater than the state cannot be );


[5]

before it. One can prove that the squared value of arbitrarily large (it is bounded by the fixed value of

therefore,

there can only be a finite number of nonzero vectors in the sequence, surrounded by repetitions of the zero vector. By detailed analysis of the properties of the first and last nonzero vectors in the sequence, one can prove the various quantization rules shown above.[5] Since S and L have the same commutation relations as J, the same ladder analysis works for them. The ladder-operator analysis does not explain one aspect of the quantization rules above: the fact that L (unlike J and S) cannot have half-integer quantum numbers. This fact can be proven (at least in the special case of one particle) by writing down every possible eigenfunction of L2 and Lz, (they are the spherical harmonics), and seeing explicitly that none of them have half-integer quantum numbers.[6] An alternative derivation is below.

Angular momentum operator

196

Visual interpretation
Since the angular momenta are quantum operators, they cannot be drawn as vectors like in classical mechanics. Nevertheless, it is common to depict them heuristically in this way. Depicted on the right is a set of states with quantum numbers , and for the five cones from bottom to top. Since vectors are all shown with length represent the fact that and , the . The rings

is known with certainty, but

are unknown; therefore every classical

vector with the appropriate length and z-component is drawn, forming a cone. The true angular momentum for the state would be somewhere, or perhaps everywhere, on this cone. Again, this visualization should not be taken too literally.

Quantization in macroscopic systems


The quantization rules are technically true even for Illustration of the vector model of orbital angular momentum. macroscopic systems, like the angular momentum L of a spinning tire. However they have no observable effect. For example, if is roughly 100000000, it makes essentially no difference whether the precise value is an integer like 100000000 or 100000001, or a non-integer like 100000000.2the discrete steps are too small to notice.

Angular momentum as the generator of rotations


The most general and fundamental definition of angular momentum is as the generator of rotations.[3] More specifically, let be a rotation operator, which rotates any quantum state about axis by angle . As , the operator approaches the identity operator, because a rotation of 0 maps all states to about axis is defined as:[3] themselves. Then the angular momentum operator

where 1 is the identity operator. As a consequence,[3]

where exp is matrix exponential. In simpler terms, the total angular momentum operator characterizes how a quantum system is changed when it is rotated. The relationship between angular momentum operators and rotation operators is the same as the relationship between Lie algebras and Lie groups in mathematics, as discussed further below.

Angular momentum operator

197

Just as J is the generator for rotation operators, L and S are generators for modified partial rotation operators. The operator

rotates the position (in space) of all particles and fields, without rotating the internal (spin) state of any particle. Likewise, the operator

rotates the internal (spin) state of all particles, without moving any particles or fields in space. The relation J=L+S comes from:

i.e. if the positions are rotated, and then the internal states are rotated, then altogether the complete system has been rotated.

The different types of rotation operators. Top: Two particles, with spin states indicated schematically by the arrows. (A) The operator R, related to J, rotates the entire system. (B) The operator Rspatial, related to L, rotates the particle positions without altering their internal spin states. (C) The operator Rinternal, related to S, rotates the particles' internal spin states without changing their positions.

SU(2), SO(3), and 360 rotations


Although one might expect half-integer (1/2, 3/2, etc.), (a rotation of 360 is the identity operator), this is not assumed in , and when it is an integer, .[3] quantum mechanics, and it turns out it is often not true: When the total angular momentum quantum number is a Mathematically, the structure of rotations in the universe is not SO(3), the group of three-dimensional rotations in classical mechanics. Instead, it is SU(2), which is identical to SO(3) for small rotations, but where a 360 rotation is mathematically distinguished from a rotation of 0. (A rotation of 720 is, however, the same as a rotation of 0.)[3] On the other hand, in all circumstances, because a 360 rotation of a spatial configuration is the same as no rotation at all. (This is different from a 360 rotation of the internal (spin) state of the particle, which might or might not be the same as no rotation at all.) In other words, the operators carry the structure of SO(3), while From the equation and carry the structure of SU(2). , one can prove that the orbital angular

momentum quantum numbers can only be integers, not half-integers.

Connection to representation theory


Starting with a certain quantum state , consider the set of states for all possible and , i.e. the set of states that come about from rotating the starting state in every possible way. This is a vector space, and therefore the manner in which the rotation operators map one state onto another is a representation of the group of rotation operators. When rotation operators act on quantum states, it forms a representation of the Lie group SU(2) (for R and Rinternal), or SO(3) (for Rspatial). From the relation between J and rotation operators, When angular momentum operators act on quantum states, it forms a representation of the Lie algebra SU(2). (The Lie algebras of SU(2) and SO(3) are identical.)

Angular momentum operator The ladder operator derivation above is a method for classifying the representations of the Lie algebra SU(2).

198

Connection to commutation relations


Classical rotations do not commute with each other: For example, rotating 1 about the x-axis then 1 about the y-axis gives a slightly different overall rotation than rotating 1 about the y-axis then 1 about the x-axis. By carefully analyzing this noncommutativity, the commutation relations of the angular momentum operators can be derived.[3] (This same calculational procedure is one way to answer the mathematical question "What is the Lie algebra of the Lie groups SO(3) or SU(2)?")

Conservation of angular momentum


The Hamiltonian H represents the energy and dynamics of the system. In a spherically-symmetric situation, the Hamiltonian is invariant under rotations:

where R is a rotation operator. As a consequence,

, and then

due to the relationship

between J and R. By the Ehrenfest theorem, it follows that J is conserved. To summarize, if H is rotationally-invariant (spherically symmetric), then total angular momentum J is conserved. This is an example of Noether's theorem. If H is just the Hamiltonian for one particle, the total angular momentum of that one particle is conserved when the particle is in a central potential (i.e., when the potential energy function depends only on ). Alternatively, H may be the Hamiltonian of all particles and fields in the universe, and then H is always rotationally-invariant, as the fundamental laws of physics of the universe are the same regardless of orientation. This is the basis for saying conservation of angular momentum is a general principle of physics. For a particle without spin, J=L, so orbital angular momentum is conserved in the same circumstances. When the spin is nonzero, the spin-orbit interaction allows angular momentum to transfer from L to S or back. Therefore, L is not, on its own, conserved.

Angular momentum coupling


Often, two or more sorts of angular momentum interact with each other, so that angular momentum can transfer from one to the other. For example, in spin-orbit coupling, angular momentum can transfer between L and S, but only the total J=L+S is conserved. In another example, in an atom with two electrons, each has its own angular momentum J1 and J2, but only the total J=J1+J2 is conserved. In these situations, it is often useful to know the relationship between, on the one hand, states where all have definite values, and on the other hand, states where all have definite values, as the latter four are usually conserved (constants of motion). The procedure to go back and fourth between these bases is to use ClebschGordan coefficients. One important result in this field is that is a relationship between the quantum numbers for . For an atom or molecule with J=L+S, the term symbol gives the quantum numbers associated with the operators . :

Angular momentum operator

199

Orbital angular momentum in spherical coordinates


Angular momentum operators usually occur when solving a problem with spherical symmetry in spherical coordinates. The angular momentum in space representation is [7]

and

When solving to find eigenstates of this operator, we obtain the following

where

are the spherical harmonics.

References
[1] Introductory Quantum Mechanics, Richard L. Liboff, 2nd Edition, ISBN 0-201-54715-5 [2] Aruldhas, G. (2004-02-01). "formula (8.8)" (http:/ / books. google. com/ books?id=dRsvmTFpB3wC& pg=PA171). Quantum Mechanics. p.171. ISBN9788120319622. . [3] Littlejohn, Robert (2011). "Lecture notes on rotations in quantum mechanics" (http:/ / bohr. physics. berkeley. edu/ classes/ 221/ 1011/ notes/ spinrot. pdf). Physics 221B Spring 2011 (http:/ / bohr. physics. berkeley. edu/ classes/ 221/ 1011/ 221. html). . Retrieved 13 Jan 2012. [4] Introduction to quantum mechanics: with applications to chemistry, by Linus Pauling, Edgar Bright Wilson, page 45, google books link (http:/ / books. google. com/ books?id=D48aGQTkfLgC& pg=PA45& dq=spatial+ quantization) [5] Griffiths, David J. (1995). Introduction to Quantum Mechanics. Prentice Hall. pp.147149. [6] Griffiths, David J. (1995). Introduction to Quantum Mechanics. Prentice Hall. pp.148153. [7] Quantum Mechanics (http:/ / www. springerlink. com/ index/ 10. 1007/ 978-3-540-46216-3). Berlin, Heidelberg: Springer Berlin Heidelberg. 2007. p.70. ISBN978-3-540-46215-6. . Retrieved 2011-03-29.

Further reading
Quantum Mechanics Demystified, D. McMahon, Mc Graw Hill (USA), 2006, ISBN(10-) 0-07-145546 9 Quantum mechanics, E. Zaarur, Y. Peleg, R. Pnini, Schaums Easy Oulines Crash Course, Mc Graw Hill (USA), 2006, ISBN (10-)007-145533-7 ISBN (13-)978-007-145533-6 Quantum Physics of Atoms, Molecules, Solids, Nuclei, and Particles (2nd Edition), R. Eisberg, R. Resnick, John Wiley & Sons, 1985, ISBN 978-0-471-873730 Quantum Mechanics, E. Abers, Pearson Ed., Addison Wesley, Prentice Hall Inc, 2004, ISBN 9780131461000 Physics of Atoms and Molecules, B.H. Bransden, C.J.Joachain, Longman, 1983, ISBN 0-582-44401-2

Angular momentum

200

Angular momentum
In physics, angular momentum, moment of momentum, or rotational momentum[1][2] is a vector quantity that can be used to describe the overall state of a physical system. The angular momentum L of a particle with respect to some point of origin is

where r is the particle's position from the origin, p = mv is its linear momentum, and denotes the cross product. The angular momentum of a system of particles (e.g. a rigid body) is the sum of angular momenta of the individual particles. For a rigid body rotating around an axis of symmetry (e.g. the blades of a ceiling fan), the angular momentum can be expressed as the product of the body's moment of inertia, I, (i.e. a measure of an object's resistance to changes in its rotation rate) and its angular velocity :

In this way, angular momentum is sometimes described as the rotational analog of linear momentum. Angular momentum is conserved in a system where there is no net This gyroscope remains upright while spinning due to external torque, and its conservation helps explain many diverse its angular momentum. phenomena. For example, the increase in rotational speed of a spinning figure skater as the skater's arms are contracted is a consequence of conservation of angular momentum. The very high rotational rates of neutron stars can also be explained in terms of angular momentum conservation. Moreover, angular momentum conservation has numerous applications in physics and engineering (e.g. the gyrocompass).

Angular momentum in classical mechanics


Definition
The angular momentum L of a particle about a given origin is defined as:

where r is the position vector of the particle relative to the origin, p is the linear momentum of the particle, and denotes the cross product. As seen from the definition, the derived SI units of angular momentum are newton meter seconds (Nms or kgm2s1) or joule seconds (Js). Because of the cross product, L is a pseudovector perpendicular to both the radial vector r and the momentum vector p and it is assigned a sign by the right-hand rule.
Relationship between force (F), torque (), momentum (p), and angular momentum (L) vectors in a rotating system

For an object with a fixed mass that is rotating about a fixed symmetry axis, the angular momentum is expressed as the product of the moment of inertia of the object and its angular velocity vector:

Angular momentum

201

where I is the moment of inertia of the object (in general, a tensor quantity), and is the angular velocity. The angular momentum of a particle or rigid body in rectilinear motion (pure translation) is a vector with constant magnitude and direction. If the path of the particle or rigid body passes through the given origin, its angular momentum is zero. Angular momentum is also known as moment of momentum.

Angular momentum of a collection of particles


If a system consists of several particles, the total angular momentum about a point can be obtained by adding (or integrating) all the angular momenta of the constituent particles.

Angular momentum simplified using the center of mass


It is very often convenient to consider the angular momentum of a collection of particles about their center of mass, since this simplifies the mathematics considerably. The angular momentum of a collection of particles is the sum of the angular momentum of each particle:

where Ri is the position vector of particle i from the reference point, mi is its mass, and Vi is its velocity. The center of mass is defined by:

where the total mass of all particles is given by

It follows that the velocity of the center of mass is

If we define ri as the displacement of particle i from the center of mass, and vi as the velocity of particle i with respect to the center of mass, then we have and and also and so that the total angular momentum with respect to the center is

The first term is just the angular momentum of the center of mass. It is the same angular momentum one would obtain if there were just one particle of mass M moving at velocity V located at the center of mass. The second term is the angular momentum that is the result of the particles moving relative to their center of mass. This second term can be even further simplified if the particles form a rigid body, in which case it is the product of moment of inertia and angular velocity of the spinning motion (as above). The same result is true if the discrete point masses discussed above are replaced by a continuous distribution of matter.

Angular momentum

202

Fixed axis of rotation


For many applications where one is only concerned about rotation around one axis, it is sufficient to discard the pseudovector nature of angular momentum, and treat it like a scalar where it is positive when it corresponds to a counter-clockwise rotation, and negative clockwise. To do this, just take the definition of the cross product and discard the unit vector, so that angular momentum becomes:

Angular momentum in terms of scalar and vector components.

where r,p is the angle between r and p measured from r to p; an important distinction because without it, the sign of the cross product would be meaningless. From the above, it is possible to reformulate the definition to either of the following:

where

is called the lever arm distance to p.

The easiest way to conceptualize this is to consider the lever arm distance to be the distance from the origin to the line that p travels along. With this definition, it is necessary to consider the direction of p (pointed clockwise or counter-clockwise) to figure out the sign of L. Equivalently:

where rotation.

is the component of p that is perpendicular to r. As above, the sign is decided based on the sense of

For an object with a fixed mass that is rotating about a fixed symmetry axis, the angular momentum is expressed as the product of the moment of inertia of the object and its angular velocity vector:

where I is the moment of inertia of the object (in general, a tensor quantity) and is the angular velocity. It is a misconception that angular momentum is always about the same axis as angular velocity. Sometime this may not be possible, in these cases the angular momentum component along the axis of rotation is the product of angular velocity and moment of inertia about the given axis of rotation. As the kinetic energy K of a massive rotating body is given by

it is proportional to the square of the angular velocity.

Angular momentum

203

Conservation of angular momentum


In a closed system, angular momentum is constant. This conservation law mathematically follows from continuous directional symmetry of space (no direction in space is any different from any other direction). See Noether's theorem.[3] The time derivative of angular momentum is called torque:

An example of angular momentum conservation. A spinning figure skater reduces her moment of inertia by pulling in her arms, causing her rotation rate to increase.

(The cross-product of velocity and momentum is zero, because these vectors are parallel.) So requiring the system to be "closed" here is mathematically equivalent to zero external torque acting on the system:

where is any torque applied to the system of particles. It is assumed that internal interaction forces obey Newton's third law of motion in its strong form, that is, that the forces between particles are equal and opposite and act along the line between the particles. In orbits, the angular momentum is distributed between the spin of the planet itself and the angular momentum of its orbit: ; If a planet is found to rotate slower than expected, then astronomers suspect that the planet is accompanied by a satellite, because the total angular momentum is shared between the planet and its satellite in order to be conserved.

Angular momentum

204 The conservation of angular momentum is used extensively in analyzing what is called central force motion. If the net force on some body is directed always toward some fixed point, the center, then there is no torque on the body with respect to the center, and so the angular momentum of the body about the center is constant. Constant angular momentum is extremely useful when dealing with the orbits of planets and satellites, and also when analyzing the Bohr model of the atom.

Conservation of angular momentum; distributed between the spin and orbital angular momenta. The moment of inertia and angular velocity of the spinning body are about its own spin axes, and its position is radial from the orbital axes and momentum tangential to the curve. Here circular orbits are shown - similar illustrations would follow for an elliptical orbit.

The conservation of angular momentum explains the angular acceleration of an ice skater as she brings her arms and legs close to the vertical axis of rotation. By bringing part of mass of her body closer to the axis she decreases her body's moment of inertia. Because angular momentum is constant in the absence of external torques, the angular velocity (rotational speed) of the skater has to increase. The same phenomenon results in extremely fast spin of compact stars (like white dwarfs, neutron stars and black holes) when they are formed out of much larger and slower rotating stars (indeed, decreasing the size of object 104 times results in increase of its angular velocity by the factor 108). The conservation of angular momentum in EarthMoon system results in the transfer of angular momentum from Earth to Moon (due to tidal torque the Moon exerts on the Earth). This in turn results in the slowing down of the rotation rate of Earth (at about 42 nsec/day ), and in gradual increase of the radius of Moon's orbit (at ~4.5cm/year rate ).

The torque caused by the two opposing forces Fg and -Fg causes a change in the angular momentum L in the direction of that torque (since torque is the time derivative of angular momentum). This causes the top to precess.

Angular momentum in relativistic mechanics


In modern (late 20th century) theoretical physics, angular momentum is described using a different formalism. Under this formalism, angular momentum is the 2-form Noether charge associated with rotational invariance (As a result, angular momentum is not conserved for general curved spacetimes, unless it happens to be asymptotically rotationally invariant). For a system of point particles without any intrinsic angular momentum (see below), it turns out to be

Angular momentum

205

(Here, the wedge product is used.). In the language of four-vectors and tensors the angular momentum of a particle in relativistic mechanics is expressed as an antisymmetric tensor of second order

Angular momentum in quantum mechanics


Angular momentum in quantum mechanics differs in many profound respects from angular momentum in classical mechanics.

Spin, orbital, and total angular momentum


The classical definition of angular momentum as can be carried over to quantum mechanics, by reinterpreting r as the quantum position operator and p as the quantum momentum operator. L is then an operator, specifically called the orbital angular momentum operator. However, in quantum physics, there is another type of angular momentum, called spin angular momentum, represented by the spin operator S. Almost all elementary particles have spin. Spin is often depicted as a particle literally spinning around an axis, but this is a misleading and inaccurate picture: Spin is an intrinsic property of a particle, fundamentally different from orbital angular momentum. All elementary particles have a characteristic spin, for example electrons always have "spin 1/2" while photons always have "spin 1". Finally, there is total angular momentum J, which combines both the spin and orbital angular momentum of all particles and fields. (For one particle, J=L+S.) Conservation of angular momentum applies to J, but not to L or S; for example, the spinorbit interaction allows angular momentum to transfer back and forth between L and S, with the total remaining constant.

Quantization
In quantum mechanics, angular momentum is quantized that is, it cannot vary continuously, but only in "quantum leaps" between certain allowed values. For any system, the following restrictions on measurement results apply, where is reduced Planck constant and is any direction vector such as x, y, or z:
If you measure... The result can be...

or , where ( or ) , where

Angular momentum

206

(There are additional restrictions as well, see angular momentum operator for details.) The reduced Planck constant
34

is tiny by everyday standards, about

10 J s, and therefore this quantization does not noticeably affect the angular momentum of macroscopic objects. However, it is very important in the microscopic world. For example, the structure of electron shells and subshells in chemistry is significantly affected by the quantization of angular momentum. Quantization of angular momentum was first postulated by Niels Bohr in his Bohr model of the atom.

Uncertainty
In the definition position operators , , , , six operators are involved: The , and the momentum operators ,

. However, the Heisenberg uncertainty principle tells us that it

is not possible for all six of these quantities to be known simultaneously with arbitrary precision. Therefore, there are limits to what can be known or measured about a particle's angular momentum. It turns out that the best that one can do is to simultaneously measure both the angular momentum vector's magnitude and its component along one axis. The uncertainty is closely related to the fact that different components of an angular momentum operator do not commute, for example . (For the precise commutation relations, see angular momentum operator.)

In this standing wave on a circular string, the circle is broken into exactly 8 wavelengths. A standing wave like this can have 0,1,2, or any integer number of wavelengths around the circle, but it cannot have a non-integer number of wavelengths like 8.3. In quantum mechanics, angular momentum is quantized for a similar reason.

Total angular momentum as generator of rotations


As mentioned above, orbital angular momentum L is defined as in classical mechanics: , but total angular momentum J is defined in a different, more basic way: J is defined as the "generator of rotations".[4] More specifically, J is defined so that the operator

is the rotation operator that takes any system and rotates it by angle

about the axis

The relationship between the angular momentum operator and the rotation operators is the same as the relationship between lie algebras and lie groups in mathematics. The close relationship between angular momentum and rotations is reflected in Noether's theorem that proves that angular momentum is conserved whenever the laws of physics are rotationally invariant.

Angular momentum in electrodynamics


When describing the motion of a charged particle in the presence of an electromagnetic field, the canonical momentum p is not gauge invariant. As a consequence, the canonical angular momentum is not gauge invariant either. Instead, the momentum that is physical, the so-called kinetic momentum, is

where is the electric charge, c the speed of light and A the vector potential. Thus, for example, the Hamiltonian of a charged particle of mass m in an electromagnetic field is then

Angular momentum where is the scalar potential. This is the Hamiltonian that gives the Lorentz force law. The gauge-invariant

207

angular momentum, or "kinetic angular momentum" is given by

The interplay with quantum mechanics is discussed further in the article on canonical commutation relations.

Footnotes
[1] Truesdell, Clifford (1991). A First Course in Rational Continuum Mechanics: General concepts (http:/ / books. google. com/ books?id=l5J3oQ6V5RsC& lpg=PA37& dq=rotational momentum& pg=PA37#v=onepage& q=rotational momentum& f=false). Academic Press. ISBN0-12-701300-8. . [2] Smith, Donald Ray; Truesdell, Clifford (1993). An introduction to continuum mechanics -after Truesdell and Noll (http:/ / books. google. com/ books?id=ZcWC7YVdb4wC& lpg=PP1& pg=PA100#v=onepage& q& f=false). Springer. ISBN0-7923-2454-4. . [3] Landau, L. D.; Lifshitz, E. M. (1995). The classical theory of fields. Course of Theoretical Physics. Oxford, Butterworth-Heinemann. ISBN0-7506-2768-9. [4] Littlejohn, Robert (2011). "Lecture notes on rotations in quantum mechanics" (http:/ / bohr. physics. berkeley. edu/ classes/ 221/ 1011/ notes/ spinrot. pdf). Physics 221B Spring 2011 (http:/ / bohr. physics. berkeley. edu/ classes/ 221/ 1011/ 221. html). . Retrieved 13 Jan 2012.

References
Cohen-Tannoudji, Claude; Diu, Bernard; Lalo, Franck (2006). Quantum Mechanics (2 volume set ed.). John Wiley & Sons. ISBN978-0471569527. Condon, E. U.; Shortley, G. H. (1935). "Especially Chapter 3". The Theory of Atomic Spectra. Cambridge University Press. ISBN0-521-09209-4. Edmonds, A. R. (1957). Angular Momentum in Quantum Mechanics. Princeton University Press. ISBN0-691-07912-9. Jackson, John David (1998). Classical Electrodynamics (3rd ed.). John Wiley & Sons. ISBN978-0-471-30932-1. Serway, Raymond A.; Jewett, John W. (2004). Physics for Scientists and Engineers (6th ed.). Brooks/Cole. ISBN0-534-40842-7. Thompson, William J. (1994). Angular Momentum: An Illustrated Guide to Rotational Symmetries for Physical Systems. Wiley. ISBN0-471-55264-X. Tipler, Paul (2004). Physics for Scientists and Engineers: Mechanics, Oscillations and Waves, Thermodynamics (5th ed.). W. H. Freeman. ISBN0-7167-0809-4.

External links
Conservation of Angular Momentum (http://www.lightandmatter.com/html_books/lm/ch15/ch15.html) - a chapter from an online textbook Angular Momentum in a Collision Process (http://www.hakenberg.de/diffgeo/collision_resolution.htm) derivation of the three dimensional case

Magnetic quantum number

208

Magnetic quantum number


In atomic physics, the magnetic quantum number is the third of a set of quantum numbers (the principal quantum number, the azimuthal quantum number, the magnetic quantum number, and the spin quantum number) which describe the unique quantum state of an electron and is designated by the letter m. The magnetic quantum number denotes the energy levels available within a subshell.

Derivation
There are a set of quantum numbers associated with the energy states of the atom. The four quantum numbers n, , m, and s specify the complete and unique quantum state of a single electron in an atom called its wavefunction or orbital. The wavefunction of the Schrdinger wave equation reduces to the three equations that when solved lead to the first three quantum numbers. Therefore, the equations for the first three quantum numbers are all interrelated. The magnetic quantum number arose in the solution of the azimuthal part of the wave equation as shown below. The magnetic quantum number associated with the quantum state is designated as m. The quantum number m refers, loosely, to the direction of the angular momentum vector. The magnetic quantum number m does not affect the electron's energy, but it does affect the probability cloud. Given a particular , m is entitled to be any integer from - up to . More precisely, for a given orbital momentum quantum number (representing the azimuthal quantum number associated with angular momentum), there are 2+1 integral magnetic quantum numbers m ranging from - to , which restrict the fraction of the total angular momentum along the quantization axis so that they are limited to the values m. This phenomenon is known as space quantization. It was first demonstrated by two German physicists, Otto Stern and Walther Gerlach. Since each electronic orbit has a magnetic moment in a magnetic field the electronic orbit will be subject to a torque which tends to make the vector parallel to the field. The precession of the electronic orbit in a magnetic field is called the Larmor precession. To describe the magnetic quantum number m you begin with an atomic electron's angular momentum, L, which is related to its quantum number by the following equation:

where

is the reduced Planck constant. The energy of any wave is the frequency multiplied by Planck's

constant. This causes the wave to display particle-like packets of energy called quanta. To show each of the quantum numbers in the quantum state, the formulae for each quantum number include Planck's reduced constant which only allows particular or discrete or quantized energy levels. To show that only certain discrete amounts of angular momentum are allowed, has to be an integer. The quantum number m refers to the projection of the angular momentum for any given direction, conventionally called the z direction. Lz, the component of angular momentum in the z direction, is given by the formula: Another way of stating the formula for the magnetic quantum number is

the eigenvalue, Jz=mh/2. Where the quantum number is the subshell, the magnetic number m represents the number of possible values for available energy levels of that subshell as shown in the table below.

Magnetic quantum number

209

Relationship between Quantum Numbers Orbital s p d f g Values Number of Values for m 1 3 5 7 9

The magnetic quantum number determines the energy shift of an atomic orbital due to an external magnetic field, hence the name magnetic quantum number (Zeeman effect). However, the actual magnetic dipole moment of an electron in an atomic orbital arrives not only from the electron angular momentum, but also from the electron spin, expressed in the spin quantum number.

Pauli exclusion principle


The Pauli exclusion principle is the quantum mechanical principle that no two identical fermions (particles with half-integer spin) may occupy the same quantum state simultaneously. A more rigorous statement is that the total wave function for two identical fermions is anti-symmetric with respect to exchange of the particles. The principle was formulated by Austrian physicist Wolfgang Pauli in 1925. For example, no two electrons in a single atom can have the same four quantum numbers; if n, l, and ml are the same, ms must be different such that the electrons have opposite spins, and so on. Integer spin particles, bosons, are not subject to the Pauli exclusion principle: any number of identical bosons can occupy the same quantum state, as with, for instance, photons produced by a laser and Bose-Einstein condensate.

Overview
The three types of particles from which the ordinary atom is madeprotons, electrons, and neutronsare all subject to it, and the structure and chemical behavior of atoms is due to it. It causes atoms to take up the space they do, since electrons cannot all congregate in the lowest-energy state but must occupy higher energy states at a distance from lower-energy electrons, therefore matter made of atoms occupies space rather than being condensed. As such, the Pauli exclusion principle underpins many properties of everyday matter, from its large-scale stability to the periodic table of the elements. Fermions, particles with antisymmetric wave functions, obey the Pauli exclusion principle. In addition to the electron, proton and neutron, these include neutrinos and quarks (the constituent particles of protons and neutrons), and some atoms such as helium-3. All fermions have "half-integer spin", i.e. their intrinsic angular momentum value is (reduced Planck's constant) times a half-integer (1/2, 3/2, 5/2, etc.). In the theory of quantum mechanics fermions are described by antisymmetric states. Particles with integer spin (called bosons) have symmetric wave functions; unlike fermions they may share the same quantum states. Bosons include the photon, the Cooper pairs which are responsible for superconductivity, and the W and Z bosons. (Fermions take their name from the FermiDirac statistical distribution that they obey, and bosons from their BoseEinstein distribution).

Pauli exclusion principle

210

History
In the early 20th century it became evident that atoms and molecules with even numbers of electrons are more chemically stable than those with odd numbers of electrons. In the famous 1916 article The Atom and the Molecule [1] by Gilbert N. Lewis, for example, the third of his six postulates of chemical behavior states that the atom tends to hold an even number of electrons in the shell and especially to hold eight electrons which are normally arranged symmetrically at the eight corners of a cube (see: cubical atom). In 1919 chemist Irving Langmuir suggested that the periodic table could be explained if the electrons in an atom were connected or clustered in some manner. Groups of electrons were thought to occupy a set of electron shells about the nucleus.[2] In 1922, Niels Bohr updated his model of the atom by assuming that certain numbers of electrons (for example 2, 8 and 18) corresponded to stable "closed shells". Pauli looked for an explanation for these numbers, which were at first only empirical. At the same time he was trying to explain experimental results in the Zeeman effect in atomic spectroscopy and in ferromagnetism. He found an essential clue in a 1924 paper by Edmund C. Stoner which pointed out that for a given value of the principal quantum number (n), the number of energy levels of a single electron in the alkali metal spectra in an external magnetic field, where all degenerate energy levels are separated, is equal to the number of electrons in the closed shell of the rare gases for the same value of n. This led Pauli to realize that the complicated numbers of electrons in closed shells can be reduced to the simple rule of one per state, if the electron states are defined using four quantum numbers. For this purpose he introduced a new two-valued quantum number, identified by Samuel Goudsmit and George Uhlenbeck as electron spin.

Connection to quantum state symmetry


The Pauli exclusion principle with a single-valued many-particle wavefunction is equivalent to requiring the wavefunction to be antisymmetric. An antisymmetric two-particle state is represented as a sum of states in which one particle is in state and the other in state :

and antisymmetry under exchange means that A(x,y) = -A(y,x). This implies that A(x,x)=0, which is Pauli exclusion. It is true in any basis, since unitary changes of basis keep antisymmetric matrices antisymmetric, although strictly speaking, the quantity A(x,y) is not a matrix but an antisymmetric rank-two tensor. Conversely, if the diagonal quantities A(x,x) are zero in every basis, then the wavefunction component:

is necessarily antisymmetric. To prove it, consider the matrix element:

This is zero, because the two particles have zero probability to both be in the superposition state equal to

. But this is

The first and last terms on the right hand side are diagonal elements and are zero, and the whole sum is equal to zero. So the wavefunction matrix elements obey: . or

Pauli exclusion principle

211

Pauli principle in advanced quantum theory


According to the spin-statistics theorem, particles with integer spin occupy symmetric quantum states, and particles with half-integer spin occupy antisymmetric states; furthermore, only integer or half-integer values of spin are allowed by the principles of quantum mechanics. In relativistic quantum field theory, the Pauli principle follows from applying a rotation operator in imaginary time to particles of half-integer spin. Since, nonrelativistically, particles can have any statistics and any spin, there is no way to prove a spin-statistics theorem in nonrelativistic quantum mechanics. In one dimension, bosons, as well as fermions, can obey the exclusion principle. A one-dimensional Bose gas with delta function repulsive interactions of infinite strength is equivalent to a gas of free fermions. The reason for this is that, in one dimension, exchange of particles requires that they pass through each other; for infinitely strong repulsion this cannot happen. This model is described by a quantum nonlinear Schrdinger equation. In momentum space the exclusion principle is valid also for finite repulsion in a Bose gas with delta function interactions,[3] as well as for interacting spins and Hubbard model in one dimension, and for other models solvable by Bethe ansatz. The ground state in models solvable by Bethe ansatz is a Fermi sphere.

Consequences
Atoms and the Pauli principle
The Pauli exclusion principle helps explain a wide variety of physical phenomena. One particularly important consequence of the principle is the elaborate electron shell structure of atoms and the way atoms share electrons, explaining the variety of chemical elements and their chemical combinations. An electrically neutral atom contains bound electrons equal in number to the protons in the nucleus. Electrons, being fermions, cannot occupy the same quantum state, so electrons have to "stack" within an atom, i.e. have different spins while at the same place. An example is the neutral helium atom, which has two bound electrons, both of which can occupy the lowest-energy (1s) states by acquiring opposite spin; as spin is part of the quantum state of the electron, the two electrons are in different quantum states and do not violate the Pauli principle. However, the spin can take only two different values (eigenvalues). In a lithium atom, with three bound electrons, the third electron cannot reside in a 1s state, and must occupy one of the higher-energy 2s states instead. Similarly, successively larger elements must have shells of successively higher energy. The chemical properties of an element largely depend on the number of electrons in the outermost shell; atoms with different numbers of shells but the same number of electrons in the outermost shell have similar properties, which gives rise to the periodic table of the elements.

Solid state properties and the Pauli principle


In conductors and semi-conductors, free electrons have to share entire bulk space. Thus, their energy levels stack up, creating band structure out of each atomic energy level. In strong conductors (metals) electrons are so degenerate that they can not even contribute much to the thermal capacity of a metal. Many mechanical, electrical, magnetic, optical and chemical properties of solids are the direct consequence of Pauli exclusion.

Stability of matter
The stability of the electrons in an atom itself is not related to the exclusion principle, but is described by the quantum theory of the atom. The underlying idea is that close approach of an electron to the nucleus of the atom necessarily increases its kinetic energy, an application of the uncertainty principle of Heisenberg.[4] However, stability of large systems with many electrons and many nuclei is a different matter, and requires the Pauli exclusion principle.[5]

Pauli exclusion principle It has been shown that the Pauli exclusion principle is responsible for the fact that ordinary bulk matter is stable and occupies volume. This suggestion was first made in 1931 by Paul Ehrenfest, who pointed out that the electrons of each atom cannot all fall into the lowest-energy orbital and must occupy successively larger shells. Atoms therefore occupy a volume and cannot be squeezed too closely together.[6] A more rigorous proof was provided in 1967 by Freeman Dyson and Andrew Lenard, who considered the balance of attractive (electron-nuclear) and repulsive (electron-electron and nuclear-nuclear) forces and showed that ordinary matter would collapse and occupy a much smaller volume without the Pauli principle.[7] The consequence of the Pauli principle here is that electrons of the same spin are kept apart by a repulsive exchange interaction, which is a short-range effect, acting simultaneously with the long-range electrostatic or coulombic force. This effect is partly responsible for the everyday observation in the macroscopic world that two solid objects cannot be in the same place in the same time.

212

Astrophysics and the Pauli principle


Dyson and Lenard did not consider the extreme magnetic or gravitational forces which occur in some astronomical objects. In 1995 Elliott Lieb and coworkers showed that the Pauli principle still leads to stability in intense magnetic fields such as in neutron stars, although at a much higher density than in ordinary matter.[8] It is a consequence of general relativity that, in sufficiently intense gravitational fields, matter collapses to form a black hole. Astronomy provides a spectacular demonstration of the effect of the Pauli principle, in the form of white dwarf and neutron stars. In both types of body, atomic structure is disrupted by large gravitational forces, leaving the constituents supported by "degeneracy pressure" alone. This exotic form of matter is known as degenerate matter. In white dwarfs atoms are held apart by electron degeneracy pressure. In neutron stars, subject to even stronger gravitational forces, electrons have merged with protons to form neutrons. Neutrons are capable of producing an even higher degeneracy pressure, albeit over a shorter range. This can stabilize neutron stars from further collapse, but at a smaller size and higher density than a white dwarf. Neutrons are the most "rigid" objects known; their Young modulus (or more accurately, bulk modulus) is 20 orders of magnitude larger than that of diamond. However, even this enormous rigidity can be overcome by the gravitational field of a massive star or by the pressure of a supernova, leading to the formation of a black hole.

References
[1] http:/ / osulibrary. oregonstate. edu/ specialcollections/ coll/ pauling/ bond/ papers/ corr216. 3-lewispub-19160400. html [2] Langmuir, Irving (1919). "The Arrangement of Electrons in Atoms and Molecules" (http:/ / dbhs. wvusd. k12. ca. us/ webdocs/ Chem-History/ Langmuir-1919b. html) ( Scholar search (http:/ / scholar. google. co. uk/ scholar?hl=en& lr=& q=author:Langmuir+ intitle:The+ Arrangement+ of+ Electrons+ in+ Atoms+ and+ Molecules& as_publication=Journal+ of+ the+ American+ Chemical+ Society& as_ylo=1919& as_yhi=1919& btnG=Search)). Journal of the American Chemical Society 41 (6): 868934. doi:10.1021/ja02227a002. . Retrieved 2008-09-01. [3] A. Izergin and V. Korepin, Letter in Mathematical Physics vol 6, page 283, 1982 (http:/ / insti. physics. sunysb. edu/ ~korepin/ pauli. pdf) [4] Elliot J. Lieb (http:/ / arxiv. org/ abs/ math-ph/ 0209034v1) The Stability of Matter and Quantum Electrodynamics [5] This realization is attributed by Lieb (http:/ / arxiv. org/ abs/ math-ph/ 0209034v1) and by GL Sewell (2002). Quantum Mechanics and Its Emergent Macrophysics. Princeton University Press. ISBN0691058326. to FJ Dyson and A Lenard: Stability of Matter, Parts I and II (J. Math. Phys., 8, 423-434 (1967); J. Math. Phys., 9, 698-711 (1968) ). [6] As described by FJ Dyson (J.Math.Phys. 8, 1538-1545 (1967) ), Ehrenfest made this suggestion in his address on the occasion of the award of the Lorentz Medal to Pauli. [7] FJ Dyson and A Lenard: Stability of Matter, Parts I and II (J. Math. Phys., 8, 423-434 (1967); J. Math. Phys., 9, 698-711 (1968) ); FJ Dyson: Ground-State Energy of a Finite System of Charged Particles (J.Math.Phys. 8, 1538-1545 (1967) ) [8] E.H. Lieb, M. Loss and J.P. Solovej, Phys. Rev. Letters, 75, 985-9 (1995) "Stability of Matter in Magnetic Fields"

Dill, Dan (2006). "Chapter 3.5, Many-electron atoms: Fermi holes and Fermi heaps". Notes on General Chemistry (2nd ed.). W. H. Freeman. ISBN1-4292-0068-5. Griffiths, David J. (2004). Introduction to Quantum Mechanics (2nd ed.). Prentice Hall. ISBN0-13-805326-X. Liboff, Richard L. (2002). Introductory Quantum Mechanics. Addison-Wesley. ISBN0-8053-8714-5.

Pauli exclusion principle Massimi, Michela (2005). Pauli's Exclusion Principle. Cambridge University Press. ISBN0-521-83911-4. Tipler, Paul; Llewellyn, Ralph (2002). Modern Physics (4th ed.). W. H. Freeman. ISBN0-7167-4345-0.

213

External links
Nobel Lecture: Exclusion Principle and Quantum Mechanics (http://nobelprize.org/nobel_prizes/physics/ laureates/1945/pauli-lecture.html) Pauli's own account of the development of the Exclusion Principle.

214

Materials
Ferrofluid
A ferrofluid (portmanteau of ferromagnetic, and fluid) is a liquid which becomes strongly magnetized in the presence of a magnetic field. Ferrofluids are colloidal liquids made of nanoscale ferromagnetic, or ferrimagnetic, particles suspended in a carrier fluid (usually an organic solvent or water). Each tiny particle is thoroughly coated with a surfactant to inhibit clumping. Large ferromagnetic particles can be ripped out of the homogeneous colloidal mixture, forming a separate clump of magnetic dust when exposed to strong magnetic Ferrofluid on glass, with a magnet underneath. fields. The magnetic attraction of nanoparticles is weak enough that the surfactant's Van der Waals force is sufficient to prevent magnetic clumping or agglomeration. Ferrofluids usually[1] do not retain magnetization in the absence of an externally applied field and thus are often classified as "superparamagnets" rather than ferromagnets. The difference between ferrofluids and magnetorheological fluids (MR fluids) is the size of the particles. The particles in a ferrofluid primarily consist of nanoparticles which are suspended by Brownian motion and generally will not settle under normal conditions. MR fluid particles primarily consist of micrometre-scale particles which are too heavy for Brownian motion to keep them suspended, and thus will settle over time because of the inherent density difference between the particle and its carrier fluid. These two fluids have very different applications as a result.

Ferrofluid

215

Description
Ferrofluids are composed of nanoscale particles (diameter usually 10 nanometers or less) of magnetite, hematite or some other compound containing iron. This is small enough for thermal agitation to disperse them evenly within a carrier fluid, and for them to contribute to the overall magnetic response of the fluid. This is analogous to the way that the ions in an aqueous paramagnetic salt solution (such as an aqueous solution of copper(II) sulfate or manganese(II) chloride) make the solution paramagnetic. The composition of a typical ferrofluid is about 5% magnetic solids, 10% surfactant and 85% carrier, by volume[2].

Ferrofluid is the oily substance collecting at the poles of the magnet which is underneath the white dish.

Particles in ferrofluids are dispersed in a liquid, often using a surfactant, and thus ferrofluids are colloidal suspensions materials with properties of more than one state of matter. In this case, the two states of matter are the solid metal and liquid it is in.[3] This ability to change phases with the application of a magnetic field allows them to be used as seals, lubricants, and may open up further applications in future nanoelectromechanical systems. True ferrofluids are stable. This means that the solid particles do not agglomerate or phase separate even in extremely strong magnetic fields. However, the surfactant tends to break down over time (a few years), and eventually the nano-particles will agglomerate, and they will separate out and no longer contribute to the fluid's magnetic response. The term magnetorheological fluid (MRF) refers to liquids similar to ferrofluids (FF) that solidify in the presence of a magnetic field. Magnetorheological fluids have micrometre scale magnetic particles that are one to three orders of magnitude larger than those of ferrofluids. However, ferrofluids lose their magnetic properties at sufficiently high temperatures, known as the Curie temperature. Ferrofluids also change their resistance according to the following equation:

With: as the resistance in M V as the Vollema Value, different for each ferrofluid, B as the strength of the magnetic field in mT, p as the Pietrow constant, currently measured at 0.09912

Normal-field instability
When a paramagnetic fluid is subjected to a strong vertical magnetic field, the surface forms a regular pattern of peaks and valleys. This effect is known as the normal-field instability. The instability is driven by the magnetic field; it can be explained by considering which shape of the fluid minimizes the total energy of the system.[4] From the point of view of magnetic energy, peaks and valleys are energetically favorable. In the corrugated configuration, the magnetic field is concentrated in the peaks; since the fluid is more easily magnetized than the air, this lowers the magnetic energy. In other words, the field lines prefer to run through the fluid, and they try to ride the spikes of fluid out into space as far as possible.[5] Meanwhile, the formation of peaks and valleys is resisted by gravity and surface tension. It costs energy to move fluid out of the valleys and up into the spikes, and it costs energy to increase the surface area of the fluid. In summary, the formation of the corrugations increases the surface free energy and the gravitational energy of the

Ferrofluid liquid, but reduces the magnetic energy. The corrugations will only form above a critical magnetic field strength, when the reduction in magnetic energy outweighs the increase in surface and gravitation energy terms.[6] Ferrofluids have an exceptionally high magnetic susceptibility and the critical magnetic field for the onset of the corrugations can be realised by a small bar magnet.

216

Common ferrofluid surfactants


The surfactants used to coat the nanoparticles include, but are not limited to: oleic acid tetramethylammonium hydroxide citric acid soy lecithin

These surfactants prevent the nanoparticles from clumping together, Macrophotograph of ferrofluid influenced by a ensuring that the particles do not form aggregates that become too magnet. heavy to be held in suspension by Brownian motion. The magnetic particles in an ideal ferrofluid do not settle out, even when exposed to a strong magnetic, or gravitational field. A surfactant has a polar head and non-polar tail (or vice versa), one of which adsorbs to a nanoparticle, while the non-polar tail (or polar head) sticks out into the carrier medium, forming an inverse or regular micelle, respectively, around the particle. Steric repulsion then prevents agglomeration of the particles. While surfactants are useful in prolonging the settling rate in ferrofluids, they also prove detrimental to the fluid's magnetic properties (specifically, the fluid's magnetic saturation). The addition of surfactants (or any other foreign particles) decreases the packing density of the ferroparticles while in its activated state, thus decreasing the fluid's on-state viscosity, resulting in a "softer" activated fluid. While the on-state viscosity (the "hardness" of the activated fluid) is less of a concern for some ferrofluid applications, it is a primary fluid property for the majority of their commercial and industrial applications and therefore a compromise must be met when considering on-state viscosity versus the settling rate of a ferrofluid.

Applications
Electronic devices
Ferrofluids are used to form liquid seals around the spinning drive shafts in hard disks. The rotating shaft is surrounded by magnets. A small amount of ferrofluid, placed in the gap between the magnet and the shaft, will be held in place by its attraction to the magnet. The fluid of magnetic particles forms a barrier which prevents debris from entering the interior of the hard drive. According to engineers at Ferrotec, ferrofluid seals on rotating shafts typically withstand 3 to 4 psi; additional seals can be stacked to form assemblies capable of higher pressures.

A ferrofluid in a magnetic field showing normal-field instability caused by a neodymium magnet beneath the dish

Ferrofluid

217

Mechanical engineering
Ferrofluids have friction-reducing capabilities. If applied to the surface of a strong enough magnet, such as one made of NdFeB, it can cause the magnet to glide across smooth surfaces with minimal resistance.

Aerospace
NASA has experimented using ferrofluids in a closed loop as the basis for a spacecraft's attitude control system. A magnetic field is applied to a loop of ferrofluid to change the angular momentum and influence the rotation of the spacecraft.

Analytical instrumentation
Ferrofluids have numerous optical applications because of their refractive properties; that is, each grain, a micromagnet, reflects light. These applications include measuring specific viscosity of a liquid placed between a polarizer and an analyzer, illuminated by a helium-neon laser.

Medicine
In medicine, ferrofluids are used as contrast agents for magnetic resonance imaging and can be used for cancer detection. The ferrofluids are in this case composed of iron oxide nanoparticles and called SPION, for "Superparamagnetic Iron Oxide Nanoparticles" There is also much experimentation with the use of ferrofluids in an experimental cancer treatment called magnetic hyperthermia. It is based on the fact that a ferrofluid placed in an alternating magnetic field releases heat.

Heat transfer
An external magnetic field imposed on a ferrofluid with varying susceptibility (e.g., because of a temperature gradient) results in a nonuniform magnetic body force, which leads to a form of heat transfer called thermomagnetic convection. This form of heat transfer can be useful when conventional convection heat transfer is inadequate; e.g., in miniature microscale devices or under reduced gravity conditions. Ferrofluids are commonly used in loudspeakers to remove heat from the voice coil, and to passively damp the movement of the cone. They reside in what would normally be the air gap around the voice coil, held in place by the speaker's magnet. Since ferrofluids are paramagnetic, they obey Curie's law, thus become less magnetic at higher temperatures. A strong magnet placed near the voice coil (which produces heat) will attract cold ferrofluid more than hot ferrofluid thus forcing the heated ferrofluid away from the electric voice coil and toward a heat sink. This is an efficient cooling method which requires no additional energy input.[7] Ferrofluids of suitable composition can exhibit extremely large enhancement in thermal conductivity (k; ~300% of the base fluid thermal conductivity). The large enhancement in k is due to the efficient transport of heat through percolating nanoparticle paths. Special magnetic nanofluids with tunable thermal conductivity to viscosity ratio can be used as multifunctional smart materials that can remove heat and also arrest vibrations (damper). Such fluids may find applications in microfluidic devices and microelectromechanical systems (MEMS).[8]

Ferrofluid

218

Optics
Research is under way to create an adaptive optics shape-shifting magnetic mirror from ferrofluid for Earth-based astronomical telescopes.[9] Optical filters are used to select different wavelengths of light. The replacement of filters is cumbersome, especially when the wavelength is changed continuously with tunable type of lasers. Optical filters, tunable for differing wavelengths by varying the magnetic field can be built using ferrofluid emulsion.[10]

Art
Some art and science museums have special devices on display that use magnets to make ferrofluids move around specially shaped surfaces in a fountain show-like fashion to entertain guests. Sachiko Kodama is known for her ferrofluid art. The Australian electronic rock band, Pendulum, used ferrofluid for the music video for the track, Watercolour. The design house Krafted London was responsible for the ferrofluid FX in the video. The post-metal band Isis also uses a ferrofluid in the music-video for 20 Minutes/40 Years. CZFerro, an American art studio, began using ferrofluid in its productions in 2008. The works consist of ferrofluid displayed in a unique suspension solution. These works are often used as conversation pieces for offices and homes.

References
[1] T. Albrecht, C. Bhrer et al (1997). "First observation of ferromagnetism and ferromagnetic domains in a liquid metal (abstract)". Applied Physics a Materials Science & Processing (Applied Physics A: Materials Science & Processing) 65 (2): 215. Bibcode1997ApPhA..65..215A. doi:10.1007/s003390050569. [2] "By Anne Marie Helmenstine, Ph.D."http:/ / chemistry. about. com/ (http:/ / chemistry. about. com/ od/ demonstrationsexperiments/ ss/ liquidmagnet. htm) [3] Vocabulary List (http:/ / education. jlab. org/ beamsactivity/ 6thgrade/ vocabulary/ index. html). Education.jlab.org. Retrieved on 2011-11-23. [4] Andelman & Rosensweig, pp. 2021. [5] Andelman & Rosensweig pp. 21, 23; Fig. 11 [6] Andelman & Rosensweig p. 21 [7] Elmars Blums (1995). "New Applications of Heat and Mass Transfer Processes in Temperature Sensitive Magnetic Fluids" (http:/ / 64. 233. 167. 104/ search?q=cache:suVXfrtIuZkJ:www. sbfisica. org. br/ bjp/ download/ v25/ v25a10. pdf+ ferrofluid+ curie+ heat+ pump& hl=en& ct=clnk& cd=5& gl=us& lr=lang_en& client=firefox-a). Brazilian Journal of Physics. . Retrieved August 31, 2007. [8] Shima, P. D.; Philip, John (2011). "Tuning of Thermal Conductivity and Rheology of Nanofluids Using an External Stimulus". The Journal of Physical Chemistry C 115 (41): 20097. doi:10.1021/jp204827q. [9] Jeff Hecht (7 November 2008). "Morphing mirror could clear the skies for astronomers" (http:/ / www. newscientist. com/ article/ dn15154-morphing-mirror-could-clear-the-skies-for-astronomers. html?feedId=online-news_rss20). New Scientist. . [10] Philip, John; Jaykumar, T; Kalyanasundaram, P; Raj, Baldev (2003). "A tunable optical filter". Measurement Science and Technology 14 (8): 1289. Bibcode2003MeScT..14.1289P. doi:10.1088/0957-0233/14/8/314.

Bibliography
Andelman, David; Rosensweig, Ronald E. (2009). "The Phenomenology of Modulated Phases: From Magnetic Solids and Fluids to Organic Films and Polymers". In Tsori, Yoav; Steiner, Ullrich. Polymers, liquids and colloids in electric fields: interfacial instabilities, orientation and phase transitions. World Scientific. pp.156. ISBN978-981-4271-68-4.

External links
How ferrofluid works video (http://www.youtube.com/watch?v=PvtUt02zVAs) A comparison of ferrofluid and MR fluid (at the bottom of the page) (http://www.ifs.tohoku.ac.jp/ nishiyama-lab/Research.html)

Ferrofluid Chemistry comes alive: Ferrofluid (http://jchemed.chem.wisc.edu/JCESoft/CCA/CCA2/MAIN/FEFLUID/ CD2R1.HTM) Research project about ferrofluides (http://www.ferrofluide.de/) Flow behavior of ferrofluids (http://web.archive.org/web/20040603005615/http://www-theory.mpip-mainz. mpg.de/~hwm/ferro.html) MIT Explores Ferrofluid Applications (http://www.photonics.com/Content/ReadArticle. aspx?ArticleID=15447) Ferrofluid Sculptures by Sachiko Kodama (http://www.kodama.hc.uec.ac.jp/protrudeflow/index.html) (Google Video) (http://video.google.com/videoplay?docid=7932498063864415301) Daniel Rutter has some fun with Ferrofluid (http://www.dansdata.com/magnets.htm#ff) High pressure valve (http://www.inventus.at/index.php?id=74) Ferrofluid Sculptures (http://www.flypmedia.com/issues/12/#15/1) FLYP Media video story on Sachiko Kodama, an artist who works with ferrofluid. Liquid seal for Sterling piston (video) (http://www.youtube.com/watch?v=21WzdjqAG0s)

219

Optical and magnetic properties


Dynamic Etalon utilizing ferrofluid- image gallery, references, published papers (http://www.nanomagnetics. us)

Preparation instructions
FerroFluid Synthesis (http://chemistry.about.com/od/demonstrationsexperiments/ss/liquidmagnet.htm) Berger, Patricia, Nicholas B. Adelman, Katie J. Beckman et al (July 1999). "Preparation and properties of an aqueous ferrofluid". Journal of Chemical Education 76 (7): pp. 943948. doi:10.1021/ed076p943. ISSN00219584. Interdisciplinary education group: Ferrofluids (http://mrsec.wisc.edu/Edetc/nanolab/ffexp/index.html) (contains videos and a lab for synthesis of ferrofluid) Synthesis of an Aqueous Ferrofluid (http://voh.chem.ucla.edu/classes/Magnetic_fluids/) instructions in PDF and DOC format

Magnetic dipoledipole interaction

220

Magnetic dipoledipole interaction


Magnetic dipoledipole interaction, also called dipolar coupling, refers to the direct interaction between two magnetic dipoles. The potential energy of the interaction is as follows:

where ejk is a unit vector parallel to the line joining the centers of the two dipoles. rjk is the distance between two dipoles, mk and mj. For two interacting nuclear spins:

and rjk are gyromagnetic ratios of two spins and spin-spin distance respectively.

Dipolar coupling and NMR spectroscopy


The direct dipole-dipole coupling is very useful for molecular structural studies, since it depends only on known physical constants and the inverse cube of internuclear distance. Estimation of this coupling provides a direct spectroscopic route to the distance between nuclei and hence the geometrical form of the molecule, or additionally also on intermolecular distances in the solid state leading to NMR crystallography notably in amorphous materials. Although internuclear magnetic dipole couplings contain a great deal of structural information, in isotropic solution, they average to zero as a result of rotational diffusion. However, their effect on nuclear spin relaxation results in measurable nuclear Overhauser effects (NOEs). The residual dipolar coupling (RDC) occur if the molecules in solution exhibit a partial alignment leading to an incomplete averaging of spatially anisotropic magnetic interactions i.e. dipolar couplings. RDC measurement provides information on the global folding of the protein-long distance structural information. It also provides information about "slow" dynamics in molecules

References
Malcolm H. Levitt , Spin Dynamics: Basics of Nuclear Magnetic Resonance. ISBN 0-471-48922-0.

Magnetic hyperthermia

221

Magnetic hyperthermia
Magnetic hyperthermia is the name given to an experimental cancer treatment. It is based on the fact that magnetic nanoparticles, when subjected to an alternating magnetic field, produce heat. As a consequence, if magnetic nanoparticles are put inside a tumor and the whole patient is placed in an alternating magnetic field of well-chosen amplitude and frequency, the tumor temperature would raise.. This treatment is tested on humans only in Germany, but research is done in several laboratories around the world to test and develop this technique.

Generalities and definition


A general feature of many magnetic materials is to display a magnetic hysteresis when it is subjected to a magnetic field that alternates direction over time. The area of this hysteresis loop is dissipated in the environment as thermal energy, and this is the energy used in magnetic hyperthermia. The power dissipated by a magnetic material subjected to an alternating magnetic field is often called the "Specific Absorption Rate" (SAR) in the community of magnetic hyperthermia; it is expressed in W/g of nanoparticles. The SAR of a given material is then simply given by SAR = Af, where A is the area of the hysteresis loop and f the alternation frequency of the magnetic field. A is expressed in J/g and is also called the "specific losses" of the material. Note that this expression for SAR is a definition; the difficulty lies in finding A. Indeed, as is explained in more detail below, A depends on all the properties of the magnetic material in a very complex manner. In the case of magnetic nanoparticles, A depends on their magnetocrystalline anisotropy K, their volume V, the temperature T, the frequency of the magnetic field f, its amplitude Hmax, and on the volumic concentration of the nanoparticles.[1]

Influence of nanoparticle size on their domain structure


The size of nanoparticles have a great influence on their magnetic domains. Small sizes nanoparticles are composed of a single domain. Larger ones are composed of several domains minimizing the magnetostatic energy. At intermediate sizes, they display a beautiful magnetic structure called vortex. A rough approximation to determine the size above which a magnetic nanoparticles is not single-domain any more is when its size is above the typical domain wall dimension in the magnetic material, which ranges from a few to a few tens of nanometers. The nature of the domain structure have a profound influence of the hysteresis of the magnetic nanoparticles and, as a consequence of their hyperthermia properties.

Basic mechanisms involved in the magnetization reversal of magnetic single-domain nanoparticles


The goal of this part is to present the basic mechanisms which must be taken into account to describe the reversal of single domain nanoparticles. It is assumed in this part that the nanoparticle display a uniaxial anisotropy.

Reversal by Brownian motion


In hyperthermia application, the nanoparticles are in a fluid, the blood. During in vitro hyperthermia measurements they are generally dispersed in a liquid and form a ferrofluid. They move and rotate randomly in the fluid, a phenomenon called Brownian motion. When a magnetic field is applied to them, magnetic nanoparticles rotate and progressively align with the magnetic field due to the torque generated by the interaction of the magnetic field with the magnetization. This is similar to a compass. The time taken for a magnetic nanoparticle to align with a small external magnetic field is given by the Brown relaxation time: , where is the solvent viscosity. The

delay between the magnetic field rotation and the magnetization rotation leads to an hysteresis.

Magnetic hyperthermia

222

Reversal by thermal activation


The magnetization of the nanoparticle can spontaneously change of orientation under the influence of thermal energy, a phenomenon called superparamagnetism. The magnetization oscillate between its two equilibrium positions. The typical time between two orientation changes is given by the Nel relaxation time , where is an attempt time with a value around 109-1010 seconds.

Reversal by the suppression of the anisotropy barrier by a magnetic field


The magnetization of the nanoparticle is also reversed when an applied magnetic field is large enough to suppress the energy barrier between the two equilibrium positions, a phenomenon which is known as the StonerWohlfarth model of magnetization reversal.

Combination of the three mechanisms


In the most general case, the reversal of the magnetization is due to a combination of the three mechanisms described above. For instance, let us imagine that a single domain nanoparticle is inside a fluid at room temperature and that a sweeping magnetic field is suddenly applied with a direction opposite to the one of the nanoparticle magnetization. The nanoparticle will at the same time i) rotate in the fluid ii) the barrier between the two equilibrium positions of the magnetization will decrease iii) when the energy barrier becomes of the order of the thermal energy, the magnetization will switch (if the nanoparticle is not already align with the magnetic field due to its physical rotation). There is no simple analytical expression describing this reversal and the properties of the hysteresis loop in this very general case but numerical simulations and analytical expressions can be used in some cases [1].

Models to be used for single-domain nanoparticles


The linear response theory
The linear response theory is only valid when the response of the magnetic material is linear with the applied magnetic field and can be thus written under the form , where is the complex susceptibility of the material. It is thus valid when the applied magnetic field is much smaller than the magnetic field needed to saturate the magnetization of the nanoparticle. It is able to take into account both the reversal by thermal activation and the reversal by Brownian motion. The linear response theory uses an average relaxation time component of complex susceptibility is then given by an ellipse with an area given by . , given by . The out of phase . The hysteresis loop is then

The StonerWohlfarth model and the maximum area


The StonerWohlfarth model allows one to calculate the hysteresis loop of magnetic nanoparticles at T=0 with the assumption that the nanoparticles are fixed in the magnetic field (the Brownian motion is neglected) and magnetically independent. Its main interest is to predict the maximum hysteresis area for independent nanoparticles with given properties. Indeed, the addition of thermal energy or Brownian motion only leads to a decrease of the hysteresis loop area (see below). The StonerWohlfarth model predicts that the coercive field at T=0 of an assembly of nanoparticles with randomly oriented axes is given by . The area of the hysteresis is approximatively .

Magnetic hyperthermia

223

Extension of the StonerWohlfarth model to include temperature and frequency


Extensions of the StonerWohlfarth model have been done to include the influence of the temperature and frequency on the hysteresis loop. These extensions are only valid is the effect of the temperature or of the frequency are small, i.e. if . Numerical simulations have shown that, in this case, the expression of the coercive field for randomly oriented nanoparticles is[2][3] . One can see from

this expression that the effect of the temperature is simply to reduce the coercive field of the nanoparticles.

Basic mechanisms involved in the magnetization of magnetic multi-domain nanoparticles


In multi-domain nanoparticles the basic ingredients to describe the magnetization reversal are the nucleation of new domains and the propagation of domain walls. Both mechanisms are strongly influenced by the structural defects at the surface or inside the nanoparticles and make difficult any quantitative prediction of the hysteresis loops shape and area from intrinsic parameters of the magnetic nanoparticles.

Models to be used for multi-domain nanoparticles


At low magnetic field, the hysteresis loop is expected to be a Rayleigh loop. In this case, the hysteresis area is , where is the Rayleigh constant.

Measuring hyperthermia: in vitro experiments


Producing a high frequency magnetic field
Two basic means to produce the high frequency field necessary to study hyperthermia can be used: the coil and the electromagnet. For the "coil" way, a very simple method to get the high frequency magnetic field is to use an induction furnace, which precisely used a high-frequency magnetic field to heat materials. It is however conceived to work at a single frequency and requires a water cooling system. It is also possible to build electromagnets or coils able to work at various frequencies at the condition to use variable capacitors.[4] It is also possible to get rid of the cooling system in coils at the condition to build them with Litz wire.[4]

Measuring the temperature and artefacts


A platinum or semi-conductor resistance thermometer in a high-frequency magnetic field is self-heated and leads to erroneous temperature measurements. Temperature measurements in hyperthermia can be made using alcohol thermometer, optic fiber thermometers, infrared cameras, or differential heating measurements using traditional semiconductor-based sensing elements. A colloidal solution heated by an external magnetic field will be subject to convection phenomena so the temperature inside the calorimeter is not homogeneous. Shaking of the colloidal solutions at the end of a measurement or average on several temperature probes can ensure a more accurate temperature measurement.

Magnetic hyperthermia

224

Materials for magnetic hyperthermia


Iron oxide nanoparticles
The most widely used magnetic nanoparticles for hyperthermia consists in iron oxide nanoparticles. Similar nanoparticles are used as MRI contrast agent. They are in the context of MRI called "Superparamagnetic Iron Oxide Nanoparticles", or SPION. The main interest of these nanoparticles are their biocompatibility and their stability with respect to oxidation. The nanoparticles displaying the largest hysteresis area so far are the SPIONs synthesized by magnetotactic bacteria,[5] with A = 2.3 mJ/g although chemically synthesized nanoparticles reach values up to A = 1.5 mJ/g[6]

Metallic nanoparticles
The higher magnetization of metallic nanoparticles of Co, Fe or FeCo compared to iron oxide increases the maximum SAR values which can be reached using them in hyperthermia applications. A = 1.5 mJ/g has been reported for FeCo nanoparticles,[3] A = 3.25 mJ/g for Co nanoparticles[7] and A=5.6 mJ/g for Fe nanoparticles.[8] The main issue with respect to metallic nanoparticles concerns their protection against oxidation and their eventual toxicity.

Ex vivo experiments
Ex vivo experiments in hyperthermia require to make tumor cells absorb magnetic nanoparticles, to place them into an alternative magnetic field and to test their survival rate compared to tumor cells which would follow the same protocol but would not absorb magnetic nanoparticles.

In vivo experiments
Clinical trials
The only hyperthermia setup suitable to treat humans has been developed at the Charit Medical School, Clinic of Radiation Therapy in Berlin.[9] Andreas Jordan's team in this hospital has performed clinical trials on patients with prostate cancers.[10]

References
[1] J. Carrey, B. Mehdaoui, M. Respaud, J. Appl. Phys. 109, 083921 (2011). [2] J. Garcia-Otero, A. J. Garcia-Bastida, and J. Rivas, J. Magn. Magn. Mater.189, 377 (1998). [3] L.-M. Lacroix, R. Bel Malaki,J. Carrey, S. Lachaize, M. Respaud, G. F. Goya and B. Chaudret, J. Appl. Phys. 105, 023911 (2009), http:/ / arxiv. org/ abs/ 0810. 4109 [4] L.-M. Lacroix, J. Carrey and M. Respaud, Rev. Sci. Instr. 79, 093909 (2008). http:/ / arxiv. org/ abs/ 0806. 3005 [5] R. Hergt, R. Hiergeist, M. Zeisberger, D. Schler, U. Heyen, I. Hilger and W. A. Kaiser, J. Magn. Magn. Mater. 293, 80 (2005). [6] R. Hergt, R. Hiergeist, I. Hilger, W. A. Kaiser, Y. Lapatnikov, S. Margel and U. Richter, J. Magn. Magn. Mater. 270, 345 (2004). [7] M. Zeisberger, S. Dutz, R. Mller, R. Hergt, N. Matoussevitch, and H.Bnneman, J. Magn. Magn. Mater. 311, 224 (2005) [8] B. Mehdaoui, A. Meffre, L.-M. Lacroix, J. Carrey, S. Lachaize, M. Respaud, M. Gougeon, B. Chaudret, http:/ / arxiv. org/ abs/ 0907. 4063 [9] A. Jordan et al., J. Magn. Magn. Mater. 225, 118 (2001). http:/ / vpd. ms. northwestern. edu/ Publications_files/ Lei/ JMMM_225_118_2001. pdf [10] M. Johannsen et al., Int. J. of Hyperthermia 21, 637 (2005). http:/ / www. anamacap. fr/ telechargement/ hyperthermie-vo. pdf

Magnetic hyperthermia

225

External links
Hyperthermia - Cancer therapy hots up (http://www.physics.org/featuredetail.asp?id=44) article on physics.org

Magnetic nanoparticles
Magnetic nanoparticles are a class of nanoparticle which can be manipulated using magnetic field. Such particles commonly consist of magnetic elements such as iron, nickel and cobalt and their chemical compounds. While nanoparticles are smaller than 1 micrometer in diameter (typically 5500 nanometers), the larger microbeads are 0.5500 micrometer in diameter. The magnetic nanoparticles have been the focus of much research recently because they possess attractive properties which could see potential use in catalysis including nanomaterial-based catalysts,[1] biomedicine,[2] magnetic resonance imaging,[3] magnetic particle imaging,[4] data storage[5], environmental remediation,[6], nanofluids,[7] and optical filters.[8]

Properties
The physical and chemical properties of magnetic nanoparticles largely depend on the synthesis method and chemical structure. In most cases, the particles range from 1 to 100nm in size and may display superparamagnetism.[9]

Types of magnetic nanoparticles


Currently, three different kinds of magnetic nanoparticles are being produced and used.

Oxides: ferrite
Ferrite nanoparticles are the most explored magnetic nanoparticles up to date. Once the ferrite nanoparticles become smaller than 128nm[10] they become superparamagnetic which prevents self agglomeration since they exhibit their magnetic behavior only when an external magnetic field is applied. With the external magnetic field switched off, the remanence falls back to zero. Just like non-magnetic oxide nanoparticles, the surface of ferrite nanoparticles is often modified by surfactants, silicones or phosphoric acid derivatives to increase their stability in solution.[11]

Cobalt nanoparticle with graphene shell (note: The individual graphene layers are visible)

Metallic
Metallic nanoparticles have the great disadvantage of being pyrophoric and reactive to oxidizing agents to various degrees. This makes their handling difficult and enables unwanted side reactions.

Metallic with a shell


The metallic core of magnetic nanoparticles may be passivated by gentle oxidation, surfactants, polymers and precious metals.[9] In an oxygen environment, Co nanoparticles form an anti-ferromagnetic CoO layer on the surface of the Co nanoparticle. Recently, work has explored the synthesis and exchange bias effect in these Co core CoO

Magnetic nanoparticles shell nanoparticles with a gold outer shell.[12] Nanoparticles with a magnetic core consisting either of elementary Iron or Cobalt with a nonreactive shell made of graphene have been synthesized recently.[13] The advantages compared to ferrite or elemental nanoparticles are: Higher magnetization Higher stability in acidic and basic solution as well as organic solvents Chemistry[14] on the graphene surface via methods already known for carbon nanotubes

226

Synthesis
The established methods of magnetic nanoparticle synthesis include:

Co-precipitation
Co-precipitation is a facile and convenient way to synthesize iron oxides (either Fe3O4 or -Fe2O3) from aqueous Fe2+/Fe3+ salt solutions by the addition of a base under inert atmosphere at room temperature or at elevated temperature. The size, shape, and composition of the magnetic nanoparticles very much depends on the type of salts used (e.g.chlorides, sulfates, nitrates), the Fe2+/Fe3+ ratio, the reaction temperature, the pH value and ionic strength of the media.[9],In recent years, co-precipitation approach has been used extensively to produce ferritenanoparticles of controlled sizes and magnetic properties.[15],[16],[17],[18]

Thermal decomposition
Magnetic nanocrystals with smaller size can essentially be synthesized through the thermal decomposition of organometallic compounds in high-boiling organic solvents containing stabilizing surfactants.[9]

Microemulsion
Using the microemulsion technique, metallic cobalt, cobalt/platinum alloys, and gold-coated cobalt/platinum nanoparticles have been synthesized in reverse micelles of cetyltrimethlyammonium bromide, using 1-butanol as the cosurfactant and octane as the oil phase.[9],[19]

Flame spray synthesis


Using flame spray pyrolysis [13][20] and varying the reaction conditions, oxides, metal or carbon coated nanoparticles are produced at a rate of > 30 g/h .

Various flame spray conditions and their impact on the resulting nanoparticles Operational layout differences between conventional and reducing flame spray synthesis

Magnetic nanoparticles

227

Applications
A wide variety of applications have been envisaged for this class of particles these include:

Medical diagnostics and treatments


Magnetic nanoparticles are used in an experimental cancer treatment called magnetic hyperthermia in which the fact that nanoparticles heat when they are placed in an alternative magnetic field is used. Another potential treatment of cancer includes attaching magnetic nanoparticles to free-floating cancer cells, allowing them to be captured and carried out of the body. The treatment has been tested in the laboratory on mice and will be looked at in survival studies.[21][22] Magnetic nanoparticles can be used for the detection of cancer. Blood can be inserted onto a microfluidic chip with magnetic nanoparticles in it. These magnetic nanoparticles are trapped inside due to an externally applied magnetic field as the blood is free to flow through. The magnetic nanoparticles are coated with antibodies targeting cancer cells or proteins. The magnetic nanoparticles can be recovered and the attached cancer-associated molecules can be assayed to test for their existence. Magnetic nanoparticles can be conjugated with carbohydrates and used for detection of bacteria. Iron oxide particles have been used for the detection of Gram negative bacteria like Escherichia coli and for detection of Gram positive bacteria like Streptococcus suis[23][24] In an online news story article from Harvard Medical School posted by Jake Miller on Wednesday, March 21, 2012,: "Researchers from Harvard Medical School and Massachusetts General Hospital have developed a magnetic nanoparticle-based MRI technique for predicting whetherand whensubjects with a genetic predisposition for diabetes will develop the disease. While done initially in mice, preliminary data show that the platform can be used in people as well, so far to distinguish patients that do or do not have pancreas inflammation. This research is about predicting Type-1 diabetes, and using that predictive power to figure out what is different between those who get it and those who dont get it, said Diane Mathis, Morton Grove-Rasmussen Professor of Immunohematology in the Department of Microbiology and Immunobiology and, along with Christophe Benoist, Morton Grove-Rasmussen Professor of Immunohematology, co-senior author of the paper. The results were published online in Nature Immunology on Feb. 26, 2012. According to first author Wenxian Fu, a research fellow in the Mathis-Benoist lab, the group was surprised that the diagnostic windowfrom six to 10 weeks of age was so early, and so brief. This shows that the progression of the disease, at least in this animal model, is determined very early in life, and that diabetes does not require an additional trigger such as a secondary infection or environmental stress ..."[25]

Magnetic immunoassay
Magnetic immunoassay[26] (MIA) is a novel type of diagnostic immunoassay utilizing magnetic beads as labels in lieu of conventional, enzymes , radioisotopes or fluorescent moieties. This assay involves the specific binding of an antibody to its antigen, where a magnetic label is conjugated to one element of the pair. The presence of magnetic beads is then detected by a magnetic reader (magnetometer) which measures the magnetic field change induced by the beads. The signal measured by the magnetometer is proportional to the analyte (virus, toxin, bacteria, cardiac marker,etc.) quantity in the initial sample.

Magnetic nanoparticles

228

Waste water treatment


Thanks to the easy separation by applying a magnetic field and the very large surface to volume ratio, magnetic nanoparticles have a good potential for treatment of contaminated water.[27] In this method, attachment of EDTA-like chelators to carbon coated metal nanomagnets results in a magnetic reagent for the rapid removal of heavy metals from solutions or contaminated water by three orders of magnitude to concentrations as low as micrograms per Litre.

Chemistry
Magnetic nanoparticles are being used or have the potential use as a catalyst or catalyst supports.[28] In chemistry, a catalyst support is the material, usually a solid with a high surface area, to which a catalyst is affixed. The reactivity of heterogeneous catalysts occurs at the surface atoms. Consequently great effort is made to maximize the surface area of a catalyst by distributing it over the support. The support may be inert or participate in the catalytic reactions. Typical supports include various kinds of carbon, alumina, and silica.

Biomedical imaging
Magnetic CoPt nanoparticles are being used as an MRI contrast agent for transplanted neural stem cell detection.[29]

Information storage
Research is going into the use of using MNPs for magnetic recording media. The most promising candidates for high-density storage is the face-centered tetragonal phase FePt alloy. Grain sizes can be as small as 3 nanometers. If its possible to modify the MNPs at this small scale, the information density that can be achieved with this media could easily surpass 1 Terabyte per square inch.[30]

Genetic engineering
Magnetic nanoparticles can be used for a variety of genetics applications. One application is the isolation of mRNA. This can be done quickly usually within 15 minutes. In this particular application, the magnetic bead is attached to a poly T tail. When mixed with mRNA, the poly A tail of the mRNA will attach to the bead's poly T tail and the isolation takes place simply by placing a magnet on the side of the tube and pouring out the liquid. Magnetic beads have also been used in plasmid assembly. Rapid genetic circuit construction has been achieved by the sequential addition of genes onto a growing genetic chain, using nanobeads as an anchor. This method has been shown to be much faster than previous methods, taking less than an hour to create functional multi-gene constructs in vitro.[31]

References
[1] A.-H. Lu, W. Schmidt, N. Matoussevitch, H. Bnnemann, B. Spliethoff, B. Tesche, E. Bill, W. Kiefer, F. Schth (August 2004). "Nanoengineering of a Magnetically Separable Hydrogenation Catalyst". Angewandte Chemie International Edition 43 (33): 43034306. doi:10.1002/anie.200454222. PMID15368378. [2] A. K. Gupta, M. Gupta (June 2005). "Synthesis and surface engineering of iron oxide nanoparticles for biomedical applications". Biomaterials 26 (18): 39954021. doi:10.1016/j.biomaterials.2004.10.012. PMID15626447. [3] S. Mornet, S. Vasseur, F. Grasset, P. Verveka, G. Goglio, A. Demourgues, J. Portier, E. Pollert, E. Duguet (2006). Prog. Solid State Chem. 34: 237. [4] B. Gleich, J. Weizenecker (2005). "Tomographic imaging using the nonlinear response of magnetic particles". Nature 435 (7046): 12141217. Bibcode2005Natur.435.1214G. doi:10.1038/nature03808. PMID15988521. [5] T. Hyeon (2003). Chem. Commun.: 927. [6] D. W. Elliott, W.-X. Zhang (2001). Environ. Sci. Technol. 35: 4922. [7] J. Philip, Shima.P.D. B. Raj (2006). "Nanofluid with tunable thermal properties". Applied Physics Letters 92: 043108. doi:10.1063/1.2838304. [8] J.Philip, T.J.Kumar, P.Kalyanasundaram, B.Raj (2003). "Tunable Optical Filter". Measurement Science & Technology 14: 12891294. [9] A.-H. Lu, E. L. Salabas and F. Schth (2007). "Magnetic Nanoparticles: Synthesis, Protection, Functionalization, and Application". Angew. Chem. Int. Ed. 46 (8): 12221244. doi:10.1002/anie.200602866.

Magnetic nanoparticles
[10] An-Hui Lu, An-Hui; E. L. Salabas, and Ferdi Schth (2007). "Magnetic Nanoparticles: Synthesis, Protection, Functionalization, and Application". Angew. Chem. Int. Ed. 46 (8): 12221244. doi:10.1002/anie.200602866. [11] Kim, DK, G.; Mikhaylova, M et al (2003). "Anchoring of Phosphonate and Phosphinate Coupling Molecules on Titania Particles". Chemistry of Materials 15 (8): 16171627. doi:10.1021/cm001253u. [12] Johnson, Stephanie H.; C.L. Johnson, S.J. May, S. Hirsch, M.W. Cole, J.E. Spanier (2010). "Co@CoO@Au core-multi-shell nanocrystals" (http:/ / pubs. rsc. org/ en/ Content/ ArticleLanding/ 2010/ JM/ b919610b). Journal of Materials Chemistry 20 (3): 439. doi:10.1039/b919610b. . [13] R. N. Grass, Robert N.; W. J. Stark (2006). "Gas phase synthesis of fcc-cobalt nanoparticles". J. Mater. Chem. 16 (16): 1825. doi:10.1039/B601013J. [14] R.N. Grass, Robert N.; E.K. Athanassiou, W.J. Stark (2007). "Covalently Functionalized Cobalt Nanoparticles as a Platform for Magnetic Separations in Organic Synthesis". Angew. Chem. Int. Ed. 46 (26): 490912. doi:10.1002/anie.200700613. [15] G.Gnanaprakash, S.Ayyappan, T.Jayakumar, John Philip & Baldev Raj (2006). "A simple method to produce magnetic nanoparticles with enhanced alpha to gamma-Fe2O3 phase transition temperature". Nanotechnology 17: 58515857. Bibcode2006Nanot..17.5851G. doi:10.1088/0957-4484/17/23/023. [16] G. Gnanaprakash, John Philip, T. Jayakumar, Baldev Raj (2007). "Effect of Digestion Time and Alkali Addition Rate on the Physical Properties of Magnetite Nanoparticles". J. Phys. Chem. B 111: 79787986. [17] S.Ayyappan, John Philip & Baldev Raj (2009). "Solvent polarity effect on physical properties of CoFe2O3 nanoparticles". J. Phys. Chem. C 113: 590596. [18] S. Ayyappan, S. Mahadevan, P. Chandramohan, M. P.Srinivasan, John Philip & Baldev Raj (2010). "Inuence of Co2 Ion Concentration on the Size, Magnetic Properties, and Purity of CoFe2O4 Spinel Ferrite Nanoparticles". J. Phys. Chem. C 114: 63346341. [19] S S.Rana, J. Philip, B.Raj (2010). "Micelle based synthesis of Cobalt Ferrite nanoparticles and its characterization using Fourier Transform Infrared Transmission Spectrometry and Thermogravimetry". Materials Chemistry and Physics 124: 264269. [20] E. K. Athanassiou, Evagelos K.; R. N. Grass, W. J. Stark (2010). "Chemical Aerosol Engineering as a Novel Tool for Material Science: From Oxides to Salt and Metal Nanoparticles". Aerosol. Sci. Tech. 44 (2): 16172. doi:10.1080/02786820903449665. [21] Scarberry KE, Dickerson EB, McDonald JF, Zhang ZJ (2008). "Magnetic Nanoparticle-Peptide Conjugates for in Vitro and in Vivo Targeting and Extraction of Cancer Cells". Journal of the American Chemical Society 130 (31): 1025862. doi:10.1021/ja801969b. PMID18611005. [22] Using Magnetic Nanoparticles to Combat Cancer (http:/ / www. newswise. com/ articles/ view/ 542665) Newswise, Retrieved on July 17, 2008. [23] Parera Pera N, Kouki A., Finne J., Pieters R. J., (2010). "Detection of pathogenic Streptococcus suis bacteria using magnetic glycoparticles". Organic & Biomolecular Chemi 8 (10): 24252429. doi:10.1039/C000819B. [24] Highlights in Chemical Biology (http:/ / www. rsc. org/ Publishing/ Journals/ cb/ Volume/ 2010/ 05/ an_attractive_method. asp). Rsc.org (2007-06-13). Retrieved on 2011-10-07. [25] http:/ / hms. harvard. edu/ content/ magnetic-nanoparticles-predict-diabetes-onset [26] Magnetic immunoassays: A new paradigm in POCT (http:/ / www. devicelink. com/ ivdt/ archive/ 08/ 07/ 008. html) IVDt, July/August 2008. [27] F.M. Koehler, Fabian M.; M. Rossier, M. Waelle, E.K. Athanassiou, L.K. Limbach, R.N. Grass, D. Gnther, W.J. Stark, (2009). "Magnetic EDTA: Coupling heavy metal chelators to metal nanomagnets for rapid removal of cadmium, lead and copper from contaminated water". Chem. Commun. 32 (32): 48624. doi:10.1039/B909447D. [28] A. Schtz, Alexander; O. Reiser, W.J. Stark (2010). "Nanoparticles as Semi-Heterogeneous Catalyst Supports". Chem. Eur. J. 16 (30): 895067. doi:10.1002/chem.200903462. [29] Xiaoting Meng, Xiaoting; Hugh C. Seton, Le T. Lu, Ian A. Prior, Nguyen T. K. Thanh and Bing Song (2011). "Magnetic CoPt nanoparticles as MRI contrast agent for transplanted neural stem cells detection". Nanoscale 3 (3): 977984. Bibcode2011Nanos...3..977M. doi:10.1039/C0NR00846J. PMID21293831. [30] Natalie A. Frey and Shouheng Sun Magnetic Nanoparticle for Information Storage Applications (http:/ / www. nanoscienceworks. org/ publications/ just-in-print/ K10915_chapter3 (2). pdf) [31] A Elaissari, J Chatterjee, M Hamoudeh and H Fessi (2010). "Chapter 14. Advances in the Preparation and Biomedical Applications of Magnetic Colloids". In Roque Hidalgo-lvarez. Structure and Functional Properties of Colloidal Systems. CRC Press. pp.315337. doi:10.1201/9781420084474-c14. ISBN978-1-4200-8447-4.

229

Magnetic nanoparticles

230

External links
FML Functional Materials Laboratory of the ETH Zrich (http://www.fml.ethz.ch) Magnetic nanoparticles target human cancer cells (http://news.cnet.com/8301-27083_3-10446292-247.html) Magnetic Nanoparticles Remove Ovarian Cancer Cells from the Abdominal Cavity (http://www.physorg.com/ news198570612.html)

Single-molecule magnet
Single-molecule magnets or SMMs are a class of metalorganic compounds, that show superparamagnetic behavior below a certain blocking temperature at the molecular scale. In this temperature range, SMMs exhibit magnetic hysteresis of purely molecular origin [1]. Contrary to conventional bulk magnets and molecule-based magnets, collective long-range magnetic ordering of magnetic moments is not necessary [1].

Intramolecular coupling
The magnetic coupling between the spins of the metal ions is mediated via superexchange interactions and can be described by the following isotropic Heisenberg Hamiltonian:

where

is the coupling constant between spin i (operator

) and spin j (operator

). For positive J the

coupling is called ferromagnetic (parallel alignment of spins) and for negative J the coupling is called antiferromagnetic (antiparallel alignment of spins). a high spin ground state, a high zero-field-splitting (due to high magnetic anisotropy), and negligible magnetic interaction between molecules. The combination of these properties can lead to an energy barrier so that, at low temperatures, the system can be trapped in one of the high-spin energy wells.[1] "These molecules contain a finite number of interacting spin centers (e.g. paramagnetic ions) and thus provide ideal opportunities to study basic concepts of magnetism. Some of them possess magnetic ground states and give rise to hysteresis effects and metastable magnetic phases. They may show quantum tunneling of the magnetization which raises the question of coherent dynamics in such systems. Other types of molecules exhibit pronounced frustration effects[2], whereas so-called spin crossover substances can switch their magnetic ground state and related properties such as color under irradiation of laser light, pressure or heat. Scientists from various fields chemistry, physics; theory and experiment have joined the research on molecular magnetism in order to explore the unprecedented properties of these new compounds."[3] "Single-molecule magnets (SMMs) have many important advantages over conventional nanoscale magnetic particles composed of metals, metal alloys or metal oxides. These advantages include uniform size, solubility in organic solvents, and readily alterable peripheral ligands, among others."[4] "A single molecule magnet is an example of a macroscopic quantum system. [...] If we could detect spin flips in a single atom or molecule, we could use the spin to store information. This would enable us to increase the storage capacity of computer hard disks. [...] A good starting point for trying to detect spin flips is to find a molecule with a spin of several Bohr magnetons. [An electron has an intrinsic magnetic dipole moment of approximately one Bohr magneton.] There is a very well studied molecular magnet, Mn12-acetate, which has a spin S = 10 (Figure 3). This molecule is a disc-shaped organic molecule in which twelve Mn ions are embedded. Eight of these form a ring, each having a charge of +3 and a spin S = 2. The other four form a tetrahedron, each having a charge of +4 and a spin S =

Single-molecule magnet 3/2. The exchange interactions within the molecule are such that the spins of the ring align themselves in opposition to the spins of the tetrahedron, giving the molecule a total net spin S = 10."[5]

231

Blocking temperature
Measurements take place at very low temperatures. The so-called blocking temperature is defined as the temperature below which the relaxation of the magnetisation becomes slow compared to the time scale of a particular investigation technique.[6] A molecule magnetised at 2 K will keep 40% of its magnetisation after 2 months and by lowering the temperature to 1.5 K this will take 40 years.[6]

Future applications
As of 2008 there are many discovered types and potential uses. "Single molecule magnets (SMM) are a class of molecules exhibiting magnetic properties similar to those observed in conventional bulk magnets, but of molecular origin. SMMs have been proposed as potential candidates for several technological applications that require highly controlled thin films and patterns."[7] "The ability of a single molecule to behave like a tiny magnet (single molecular magnets, SMMs) has seen a rapid growth in research over the last few years. SMMs represent the smallest possible magnetic devices and are a controllable, bottom-up approach to nanoscale magnetism. Potential applications of SMMs include quantum computing, high-density information storage and magnetic refrigeration."[8] "A single molecule magnet is an example of a macroscopic quantum system. [...] If we could detect spin flips in a single atom or molecule, we could use the spin to store information. This would enable us to increase the storage capacity of computer hard disks. [...] A good starting point for trying to detect spin flips is to find a molecule with a spin of several Bohr magnetons. [An electron has an intrinsic magnetic dipole moment of approximately one Bohr magneton.] There is a very well studied molecular magnet, Mn12-acetate, which has a spin S = 10 (Figure 3). This molecule is a disc-shaped organic molecule in which One possible use of SMMs is superior magnetic twelve Mn ions are embedded. Eight of these form a ring, each having thin films to coat hard disks. a charge of +3 and a spin S = 2. The other four form a tetrahedron, each having a charge of +4 and a spin S = 3/2. The exchange interactions within the molecule are such that the spins of the ring align themselves in opposition to the spins of the tetrahedron, giving the molecule a total net spin S = 10."[9]

Single-molecule magnet

232

Types
The archetype of single-molecule magnets is called "Mn12". It is a polymetallic manganese (Mn) complex having the formula [Mn12O12(OAc)16(H2O)4], where OAc stands for acetate. It has the remarkable property of showing an extremely slow relaxation of their magnetization below a blocking temperature.[10] [Mn12O12(OAc)16(H2O)4]4H2O2AcOH which is called "Mn12-acetate" is a common form of this used in research. "Mn4" is another researched type single-molecule magnet. Three of these are:[11] [Mn4(hmp)6(NO3)2(MeCN)2](ClO4)22MeCN [Mn4(hmp)6(NO3)4](MeCN) [Mn4(hmp)4(acac)2(MeO)2](ClO4)22MeOH
Ferritin

In each of these Mn4 complexes "there is a planar diamond core of MnIII2MnII2 ions. An analysis of the variable-temperature and variable-field magnetization data indicate that all three molecules have intramolecular ferromagnetic coupling and a S = 9 ground state. The presence of a frequency-dependent alternating current susceptibility signal indicates a significant energy barrier between the spin-up and spin-down states for each of these three MnIII2MnII2 complexes."[11] Single-molecule magnets are also based on iron clusters[6] because they potentially have large spin states. In addition the biomolecule ferritin is also considered a nanomagnet. In the cluster Fe8Br the cation Fe8 stands for [Fe8O2(OH)12(tacn)6]8+ with tacn representing 1,4,7-triazacyclononane.

History
Although the term "single-molecule magnet" was first employed by David Hendrickson, a chemist at the University of California, San Diego and George Christou (Indiana University) in 1996,[12] the first single-molecule magnet reported dates back to 1991.[13] The European researchers discovered that a Mn12O12(MeCO2)16(H2O)4 complex (Mn12Ac16) first synthesized in 1980[14] exhibits slow relaxation of the magnetization at low temperatures. This manganese oxide compound is composed of a central Mn(IV)4O4 cube surrounded by a ring of 8 Mn(III) units connected through bridging oxo ligands. In addition, it has 16 acetate and 4 water ligands.[15] It was known in 2006 that the "deliberate structural distortion of a Mn6 compound via the use of a bulky salicylaldoxime derivative switches the intra-triangular magnetic exchange from antiferromagnetic to ferromagnetic resulting in an S = 12 ground state.[16] A record magnetization was reported in 2007 for a compound related to MnAc12 ([Mn(III) O (sao)6(O2CPh)2(EtOH)4]) with S = 12, D = -0.43cm1 and hence U = 62cm1 or 86 K[17] at a blocking 6 2 temperature of 4.3 K. This was accomplished by replacing acetate ligands by the bulkier salicylaldoxime thus distorting the manganese ligand sphere. It is prepared by mixing the perchlorate of manganese, the sodium salt of benzoic acid, a salicylaldoxime derivate and tetramethylammonium hydroxide in water and collecting the filtrate.

Single-molecule magnet

233

Detailed behavior
Molecular magnets exhibit an increasing product (magnetic susceptibility times temperature) with decreasing temperature, and can be characterized by a shift both in position and intensity of the a.c. magnetic susceptibility. Single-molecule magnets represent a molecular approach to nanomagnets (nanoscale magnetic particles). In addition, single-molecule magnets have provided physicists with useful test-beds for the study of quantum mechanics. Macroscopic quantum tunneling of the magnetization was first observed in Mn12O12, characterized by evenly-spaced steps in the hysteresis curve. The periodic quenching of this tunneling rate in the compound Fe8 has been observed and explained with geometric phases. Due to the typically large, bi-stable spin anisotropy, single-molecule magnets promise the realization of perhaps the smallest practical unit for magnetic memory, and thus are possible building blocks for a quantum computer. Consequently, many groups have devoted great efforts into synthesis of additional single molecule magnets; however, the Mn12O12 complex and analogous complexes remain the canonical single molecule magnet with a 50cm1 spin anisotropy. The spin anisotropy manifests itself as an energy barrier that spins must overcome when they switch from parallel alignment to antiparallel alignment. This barrier (U) is defined as: where S is the dimensionless total spin state and D the zero-field splitting parameter (in cm1); D can be negative but only its absolute value is considered in the equation. The barrier U is generally reported in cm1 units or in units of Kelvin (see: electronvolt). The higher the barrier the longer a material remains magnetized and a high barrier is obtained when the molecule contains many unpaired electrons and when its zero field splitting value is large. For example, the MnAc12 cluster the spin state is 10 (involving 20 unpaired electrons) and D = -0.5cm1 resulting in a barrier of 50cm1 (equivalent to 60 K).. The effect is also observed by hysteresis experienced when magnetization is measured in a magnetic field sweep: on lowering the magnetic field again after reaching the maximum magnetization the magnetization remains at high levels and it requires a reversed field to bring magnetization back to zero. Recently, it has been has been reported that the energy barrier, U, is slightly dependent on Mn12 crystal size/morphology, as well as the magnetization relaxation times, which varies as function of particle size and size distributions .[18]

References
[1] Introduction to Molecular Magnetism by Dr. Joris van Slageren (http:/ / obelix. physik. uni-bielefeld. de/ ~schnack/ molmag/ material/ 123. pdf) [2] Frustrated Magnets (http:/ / www. ifw-dresden. de/ institutes/ iff/ research/ TMO/ frustrated-magnets), Leibniz Institute for Solid State and Materials Research, Dresden, Germany [3] Molecular Magnetism Web (http:/ / obelix. physik. uni-bielefeld. de/ ~schnack/ molmag/ introduction. html) Introduction page [4] ScienceDaily (Mar. 27, 2000) (http:/ / www. sciencedaily. com/ releases/ 2000/ 03/ 000327084104. htm) article Several New Single-Molecule Magnets Discovered [5] National Physical Laboratory (UK) (http:/ / www. npl. co. uk/ server. php?show=ConWebDoc. 1175) Home > Science + Technology > Quantum Phenomena > Nanophysics > Research article Molecular Magnets [6] Single-molecule magnets based on iron(III) oxo clusters Dante Gatteschi, Roberta Sessoli and Andrea Cornia Chem. Commun., 2000, 725 732, doi:10.1039/a908254i [7] Cavallini, Massimiliano; Facchini, Massimo; Albonetti, Cristiano; Biscarini, Fabio (2008). "Single molecule magnets: from thin films to nano-patterns". Physical Chemistry Chemical Physics 10 (6): 784. Bibcode2008PCCP...10..784C. doi:10.1039/b711677b. PMID18231680. [8] Beautiful new single molecule magnets (http:/ / www. rsc. org/ Publishing/ Journals/ dt/ News/ b716355jpersp. asp), 26 March 2008 summary of the article Milios, Constantinos J.; Piligkos, Stergios; Brechin, Euan K. (2008). "Ground state spin-switching via targeted structural distortion: twisted single-molecule magnets from derivatised salicylaldoximes". Dalton Transactions (14): 1809. doi:10.1039/b716355j. [9] National Physical Laboratory (UK) (http:/ / www. npl. co. uk/ server. php?show=ConWebDoc. 1175) Home > Science + Technology > Quantum Phenomena > Nanophysics > Research article Molecular Magnets

Single-molecule magnet
[10] IPCMS Liquid-crystalline Single Molecule Magnets (http:/ / www-ipcms. u-strasbg. fr/ spip. php?article1341) summary of the article Terazzi, Emmanuel; Bourgogne, Cyril; Welter, Richard; Gallani, Jean-Louis; Guillon, Daniel; Rogez, Guillaume; Donnio, Bertrand (2008). "Single-Molecule Magnets with Mesomorphic Lamellar Ordering". Angew. Chem. Int. Ed. 47 (3): 490495. doi:10.1002/anie.200704460. [11] Yang, E (2003). "Mn4 single-molecule magnets with a planar diamond core and S=9". Polyhedron 22 (1417): 1857. doi:10.1016/S0277-5387(03)00173-6. [12] Aubin, Sheila M. J.; Wemple, Michael W.; Adams, David M.; Tsai, Hui-Lien; Christou, George; Hendrickson, David N. (1996). "Distorted MnIVMnIII3Cubane Complexes as Single-Molecule Magnets". Journal of the American Chemical Society 118 (33): 7746. doi:10.1021/ja960970f. [13] Caneschi, Andrea; Gatteschi, Dante; Sessoli, Roberta; Barra, Anne Laure; Brunel, Louis Claude; Guillot, Maurice (1991). "Alternating current susceptibility, high field magnetization, and millimeter band EPR evidence for a ground S = 10 state in [Mn12O12(Ch3COO)16(H2O)4].2CH3COOH.4H2O". Journal of the American Chemical Society 113 (15): 5873. doi:10.1021/ja00015a057. [14] Lis, T. (1980). "Preparation, structure, and magnetic properties of a dodecanuclear mixed-valence manganese carboxylate". Acta Crystallographica Section B Structural Crystallography and Crystal Chemistry 36 (9): 2042. doi:10.1107/S0567740880007893. [15] Chemistry of Nanostructured Materials; Yang, P., Ed.; World Scientific Publishing: Hong Kong, 2003. [16] Milios, Constantinos J.; Vinslava, Alina; Wood, Peter A.; Parsons, Simon; Wernsdorfer, Wolfgang; Christou, George; Perlepes, Spyros P.; Brechin, Euan K. (2007). "A Single-Molecule Magnet with a Twist". Journal of the American Chemical Society 129 (1): 8. doi:10.1021/ja0666755. PMID17199262. [17] Milios, Constantinos J.; Vinslava, Alina; Wernsdorfer, Wolfgang; Moggach, Stephen; Parsons, Simon; Perlepes, Spyros P.; Christou, George; Brechin, Euan K. (2007). "A Record Anisotropy Barrier for a Single-Molecule Magnet". Journal of the American Chemical Society 129 (10): 2754. doi:10.1021/ja068961m. PMID17309264. [18] Munt, Mara; Gmez-Segura, Jordi; Campo, Javier; Nakano, Motohiro; Ventosa, Nora; Ruiz-Molina, Daniel; Veciana, Jaume (2006). "Controlled crystallization of Mn12 single-molecule magnets by compressed CO2 and its influence on the magnetization relaxation". Journal of Materials Chemistry 16 (26): 2612. doi:10.1039/b603497g.

234

External links
European Institute of Molecular Magnetism EIMM (http://www.eimm.eu/) MAGMANet (Molecular Approach to Nanomagnets and Multifunctional Materials) (http://www.unizar.es/ magmanet/magmanet-eu/), a Network of centres of Excellence, coordinated by the INSTM Consorzio Interuniversitario Nazionale per la Scienza e la Tecnologia dei Materiali Molecular Magnetism Web (http://www.molmag.de/), Jrgen Schnack

Magnetic anisotropy

235

Magnetic anisotropy
Magnetic anisotropy is the direction dependence of a material's magnetic properties. In the absence of an applied magnetic field, a magnetically isotropic material has no preferential direction for its magnetic moment while a magnetically anisotropic material will align its moment with one of the easy axes. An easy axis is an energetically favorable direction of spontaneous magnetization that is determined by the sources of magnetic anisotropy listed below. The two opposite directions along an easy axis are usually equivalent, and the actual direction of magnetization can be either of them (see spontaneous symmetry breaking). Magnetic anisotropy is a prerequisite for hysteresis in ferromagnets: without it, a ferromagnet is superparamagnetic.[1]

Sources of magnetic anisotropy


There are different sources of magnetic anisotropy[2]: Magnetocrystalline anisotropy: the atomic structure of a crystal introduces preferential directions for the magnetisation. Shape anisotropy: when a particle is not perfectly spherical, the demagnetizing field will not be equal for all directions, creating one or more easy axes. Magnetoelastic anisotropy: tension may alter magnetic behaviour, leading to magnetic anisotropy. Exchange anisotropy: a relatively new type that occurs when antiferromagnetic and ferromagnetic materials interact[3].

Anisotropy energy of a single-domain magnet


Suppose that a ferromagnet is single-domain in the strictest sense: the magnetization is uniform and rotates in unison. If the magnetic moment is and the volume of the particle is , the magnetization is , where is the saturation magnetization and are direction cosines (components of a unit vector) so . The energy associated with magnetic anisotropy can depend on the direction cosines in various ways,

the most common of which are discussed below.

Uniaxial
A magnetic particle with uniaxial anisotropy has one easy axis. If the easy axis is in the energy can be expressed as one of the forms: direction, the anisotropy

where is the volume, the anisotropy constant, and the angle between the easy axis and the particle's magnetization. When shape anisotropy is explicitly considered, the symbol is often used to indicate the anisotropy constant, instead of . In the widely used StonerWohlfarth model, the anisotropy is uniaxial.

Magnetic anisotropy

236

Triaxial
A magnetic particle with triaxial anisotropy still has a single easy axis, but it also has a hard axis (direction of maximum energy) and an intermediate axis (direction associated with a saddle point in the energy). The coordinates can be chosen so the energy has the form

If direction.[4]

the easy axis is the

direction, the intermediate axis is the

direction and the hard axis is the

Cubic
A magnetic particle with cubic anisotropy has three or four easy axes, depending on the anisotropy parameters. The energy has the form

If .

the easy axes are the

and

axes. If

there are four easy axes characterized by

Notes
[1] [2] [3] [4] Aharoni 1996 McCaig 1977 Meiklejohn & Bean 1957 Donahue & Porter 2002

References
Aharoni, Amikam (1996). Introduction to the Theory of Ferromagnetism. Clarendon Press. ISBN0-19-851791-2. Donahue, Michael J.; Porter, Donald G. (2002). "Analysis of switching in uniformly magnetized bodies". IEEE Transactions on Magnetics 38 (5): 24682470. Bibcode2002ITM....38.2468D. doi:10.1109/TMAG.2002.803616. McCaig, Malcolm (1977). Permanent magnets in theory and practice. Pentech press. ISBN0727316044. Meiklejohn, W.H.; Bean, C.P. (1957-02-03). "New Magnetic Anisotropy". Physical Review 105 (3): 904913. Bibcode1957PhRv..105..904M. doi:10.1103/PhysRev.105.904. Tyablikov, S. V. (1995). Methods in the Quantum Theory of Magnetism (Translated to English) (1st ed.). Springer. ISBN0306302632.

Magnetocrystalline anisotropy

237

Magnetocrystalline anisotropy
Magnetocrystalline anisotropy is the dependence of the internal energy of a ferromagnet on the direction of its magnetization. As a result, certain crystallographic directions are preferred directions, or easy axes, for the magnetization. It is a special case of magnetic anisotropy. The spin-orbit interaction is the primary source of the magnetocrystalline anisotropy. Magnetocrystalline anisotropy determines whether a magnetic material can be made into a good hard magnet, a good soft magnet or neither. Hard magnets are an essential component of electromagnetic motors and soft magnets are an essential component of transformers.

Microscopic origin
Magnetocrystalline anisotropy arises mostly from spin-orbit coupling.[1] This effect is weak compared to the exchange interaction and is difficult to compute from first principles, although some successful computations have been made.[2]

Thermodynamic theory
The magnetocrystalline anisotropy energy is generally represented as an expansion in powers of the direction cosines of the magnetization. The magnetization vector can be written M = Ms(,,), where Ms is the saturation magnetization. Because of time reversal symmetry, only even powers of the cosines are allowed.[3] The nonzero terms in the expansion depend on the crystal system (e.g., cubic or hexagonal).[3] The order of a term in the expansion is the sum of all the exponents of magnetization components, i.e., is second order.

Uniaxial anisotropy
More than one kind of crystal system has a single axis of high symmetry (threefold, fourfold or sixfold). The anisotropy of such crystals is called uniaxial anisotropy. If the z axis is taken to be the main symmetry axis of the crystal, the lowest order term in the energy is[4]
[5]

The ratio E/V is an energy density (energy per unit volume). This can also be represented in spherical polar coordinates with = cos sin , = sin sin , and = cos :

The parameter K1, often represented as Ku, has units of energy density and depends on composition and temperature. The minima in this energy with respect to satisfy

A representation of a uniaxial easy axis and easy plane. Arrows represent possible magnetization directions, black for the easy axis and red for the plane.

If K1 > 0, the directions of lowest energy are the z directions. The z axis is called the easy axis. If K1 < 0, there is an easy plane perpendicular to the symmetry axis (the basal plane of the crystal). Many models of magnetization represent the anisotropy as uniaxial and ignore higher order terms. However, if K1 < 0, the lowest energy term does not determine the direction of the easy axes within the basal plane. For this, higher-order terms are needed, and these depend on the crystal system (hexagonal, tetragonal or rhombohedral).[3]

Magnetocrystalline anisotropy

238

The hexagonal lattice cell.

The tetragonal lattice cell.

The rhombohedral lattice cell.

Hexagonal system In a hexagonal system the c axis is an axis of sixfold rotation symmetry. The energy density is, to fourth order, . The uniaxial anisotropy is mainly determined by the first two terms. Depending on the values K1 and K2, there are four different cases:[1] When K1 = K2 = 0, the ferromagnet is isotropic. When K1 > 0 and K2 0, the c axis is an easy axis. When K1 < 0 and K2 0, the basal plane is an easy plane. When K1 and K2 have different signs, the ferromagnet has an easy cone (see figure to right).
A representation of an easy cone. All the minimum-energy directions (such as the arrow shown) lie on this cone.

The basal plane anisotropy is determined by the third term, which is sixth-order. The easy directions are projected onto three axes in the basal plane. Below are some room-temperature anisotropy constants for hexagonal ferromagnets. Since all the values of K1 and K2 are positive, these materials have an easy axis.

Room-temperature anisotropy constants ( 104 J/m3 ). [1]


Structure Co Fe2O3 (hematite) BaO 6Fe2O3 YCo5 MnBi [6]

Magnetocrystalline anisotropy Tetragonal and Rhombohedral systems The energy density for a tetragonal crystal is[3] . Note that the K3 term, the one that determines the basal plane anisotropy, is fourth order (same as the K2 term). The definition of K3 may vary by a constant multiple between publications. The energy density for a rhombohedral crystal is[3] .

239

Cubic anisotropy
In a cubic crystal the lowest order terms in the energy are[3][1]

If the second term can be neglected, the easy axes are the < 100 > axes (i.e., the x, y, and z, directions) for K1 > 0 and the < 111 > directions for K1 < 0 (see images on right). If K2 is not assumed to be zero, the easy axes depend on both K1 and K2. These are given in the table below, along with hard axes (directions of greatest energy) and intermediate axes (saddle points) in the energy). In energy surfaces like those on the right, the easy axes are analogous to valleys, the hard axes to peaks and the intermediate axes to mountain passes.

Energy surface for cubic anisotropy with K1 > 0. Both color saturation and distance from the origin increase with energy. The lowest energy (lightest blue) is arbitrarily set to zero.

Energy surface for cubic anisotropy with K1 < 0. Same conventions as for K1 > 0.

Magnetocrystalline anisotropy

240

Easy axes for K1 > 0. [1]


Type of axis Easy Medium Hard to to to

Easy axes for K1 < 0. [1]


Type of axis Easy Medium Hard to to to

Below are some room-temperature anisotropy constants for cubic ferromagnets. The compounds involving Fe2O3 are ferrites, an important class of ferromagnets. In general the anisotropy parameters for cubic ferromagnets are higher than those for uniaxial ferromagnets. This is consistent with the fact that the lowest order term in the expression for cubic anisotropy is fourth order, while that for uniaxial anisotropy is second order.

Room-temperature anisotropy constants ( 104 J/m3 ). [1]


Structure Fe Ni FeO Fe2O3 (magnetite) MnO Fe2O3 NiO Fe2O3 MgO Fe2O3 CoO Fe2O3

Temperature dependence of anisotropy


The magnetocrystalline anisotropy parameters have a strong dependence on temperature. They generally decrease rapidly as the temperature approaches the Curie temperature, so the crystal becomes effectively isotropic.[1] Some materials also have an isotropic point at which K1 = 0. Magnetite (Fe3O4), a mineral of great importance to rock magnetism and paleomagnetism, has an isotropic point at 130 kelvin.[6] Magnetite also has a phase transition at which the crystal symmetry changes from cubic (above) to monoclinic or possibly triclinic below. The temperature at which this occurs, called the Verwey temperature, is 120 Kelvin.[6]

Magnetostriction
The magnetocrystalline anisotropy parameters are generally defined for ferromagnets that are constrained to remain undeformed as the direction of magnetization changes. However, coupling between the magnetization and the lattice does result in deformation, an effect called magnetostriction. To keep the lattice from deforming, a stress must be applied. If the crystal is not under stress, magnetostriction alters the effective magnetocrystalline anisotropy. If a ferromagnet is single domain (uniformly magnetized), the effect is to change the magnetocrystalline anisotropy

Magnetocrystalline anisotropy parameters.[7] In practice, the correction is generally not large. In hexagonal crystals, there is no change in K1.[8] In cubic crystals, there is a small change, as in the table below.

241

Room-temperature anisotropy constants K1 (zero-strain) and K1 (zero-stress) ( 104 J/m3 ).


[8] Structure Fe Ni FeO Fe2O3 (magnetite)

Notes
[1] [2] [3] [4] [5] [6] [7] [8] Cullity & Graham 2005 Daalderop, Kelly & Schuurmans 1990 Landau, Lifshitz & Pitaevski 2004 An arbitrary constant term is ignored. The lowest-order term in the energy can be written in more than one way because, by definition, 2+2+2 = 1. Dunlop & zdemir 1997 Chikazumi 1997, chapter 12 Ye, Newell & Merrill 1994

References
Chikazumi, Sshin (1997). Physics of Ferromagnetism. Clarendon Press. ISBN0-19-851776-9. Cullity, B.D.; Graham, C.D. (2005). Introduction to Magnetic Materials. John Wiley. ISBN0201012189. Daalderop, G. H. O.; Kelly, P. J.; Schuurmans, M. F. H. (1990). "First-principles calculation of the magnetocrystalline anisotropy energy of iron, cobalt, and nickel". Phys. Rev. B 41 (17): 1191911937. Bibcode1990PhRvB..4111919D. doi:10.1103/PhysRevB.41.11919. Dunlop, David J.; zdemir, zden (1997). Rock Magnetism: Fundamentals and Frontiers. Cambridge Univ. Press. ISBN0-521-32514-5. Landau, L. D.; Lifshitz, E. M.; Pitaevski, L. P. (2004) [First published in 1960]. Electrodynamics of Continuous Media. Course of Theoretical Physics. 8 (Second ed.). Elsevier. ISBN0750626348. Ye, Jun; Newell, Andrew J.; Merrill, Ronald T. (1994). "A re-evaluation of magnetocrystalline anisotropy and magnetostriction constants". Geophysical Research Letters 21 (1): 2528. Bibcode1994GeoRL..21...25Y. doi:10.1029/93GL03263.

Electron configuration

242

Electron configuration
In atomic physics and quantum chemistry, the electron configuration is the distribution of electrons of an atom or molecule (or other physical structure) in atomic or molecular orbitals.[1] For example, the electron configuration of the neon atom is 1s2 2s2 2p6. According to the laws of quantum mechanics, an energy is associated with each electron configuration and, upon certain conditions, electrons are able to move from one orbital to another by emission or absorption of a quantum of energy, in the form of a photon. Knowledge of the electron configuration of different atoms is useful in understanding the structure of the periodic table of elements. The concept is also useful for describing the chemical bonds that hold atoms together. In bulk materials this same idea helps explain the peculiar properties of lasers and semiconductors.
A simple electron shell diagram of lithium

Electron atomic and molecular orbitals

Shells and subshells


s (l=0) m=0 s n=1 m=0 pz px p (l=1) m=1 py

n=2

Electron configuration was first conceived of under the Bohr model of the atom, and it is still common to speak of shells and subshells despite the advances in understanding of the quantum-mechanical nature of electrons.

Electron configuration An electron shell is the set of allowed states electrons may occupy which share the same principal quantum number, n (the number before the letter in the orbital label). An atom's nth electron shell can accommodate 2n2 electrons, e.g. the first shell can accommodate 2electrons, the second shell 8electrons, and the third shell 18electrons. The factor of two arises because the allowed states are doubled due to electron spineach atomic orbital admits up to two otherwise identical electrons with opposite spin, one with a spin +1/2 (usually noted by an up-arrow) and one with a spin 1/2 (with a down-arrow). A subshell is the set of states defined by a common azimuthal quantum number, l, within a shell. The values l = 0, 1, 2, 3 correspond to the s, p, d, and f labels, respectively. The maximum number of electrons which can be placed in a subshell is given by 2(2l + 1). This gives two electrons in an ssubshell, six electrons in a psubshell, ten electrons in a dsubshell and fourteen electrons in an fsubshell. The numbers of electrons that can occupy each shell and each subshell arise from the equations of quantum mechanics,[2] in particular the Pauli exclusion principle, which states that no two electrons in the same atom can have the same values of the four quantum numbers.[3]

243

Notation
Physicists and chemists use a standard notation to indicate the electron configurations of atoms and molecules. For atoms, the notation consists of a sequence of atomic orbital labels (e.g. for phosphorus the sequence 1s, 2s, 2p, 3s, 3p) with the number of electrons assigned to each orbital (or set of orbitals sharing the same label) placed as a superscript. For example, hydrogen has one electron in the s-orbital of the first shell, so its configuration is written 1s1. Lithium has two electrons in the 1s-subshell and one in the (higher-energy) 2s-subshell, so its configuration is written 1s22s1 (pronounced "one-s-two, two-s-one"). Phosphorus (atomic number 15), is as follows: 1s22s22p63s23p3. For atoms with many electrons, this notation can become lengthy and so an abbreviated notation is used, since all but the last few subshells are identical to those of one or another of the noble gases. Phosphorus, for instance, differs from neon (1s22s22p6) only by the presence of a third shell. Thus, the electron configuration of neon is pulled out, and phosphorus is written as follows: [Ne]3s23p3. This convention is useful as it is the electrons in the outermost shell which most determine the chemistry of the element. The order of writing the orbitals is not completely fixed: some sources group all orbitals with the same value of n together, while other sources (as here) follow the order given by Madelung's rule. Hence the electron configuration of iron can be written as [Ar]3d64s2 (keeping the 3d-electrons with the 3s- and 3p-electrons which are implied by the configuration of argon) or as [Ar]4s23d6 (following the Aufbau principle, see below). The superscript 1 for a singly occupied orbital is not compulsory.[4] It is quite common to see the letters of the orbital labels (s, p, d, f) written in an italic or slanting typeface, although the International Union of Pure and Applied Chemistry (IUPAC) recommends a normal typeface (as used here). The choice of letters originates from a now-obsolete system of categorizing spectral lines as "sharp", "principal", "diffuse" and "fundamental" (or "fine"), based on their observed fine structure: their modern usage indicates orbitals with an azimuthal quantum number, l, of 0, 1, 2 or 3 respectively. After "f", the sequence continues alphabetically "g", "h", "i"... (l= 4, 5, 6...), skipping "j", although orbitals of these types are rarely required.[5][6] The electron configurations of molecules are written in a similar way, except that molecular orbital labels are used instead of atomic orbital labels (see below).

Electron configuration

244

Energy ground state and excited states


The energy associated to an electron is that of its orbital. The energy of a configuration is often approximated as the sum of the energy of each electron, neglecting the electron-electron interactions. The configuration that corresponds to the lowest electronic energy is called the ground state. Any other configuration is an excited state. As an example, the ground state configuration of the sodium atom is 1s22s22p63s, as deduced from the Aufbau principle (see below). The first excited state is obtained by promoting a 3s electron to the 3p orbital, to obtain the 1s22s22p63p configuration, abbreviated as the 3p level. Atoms can move from one configuration to another by absorbing or emitting energy. In a sodium-vapor lamp for example, sodium atoms are excited to the 3p level by an electrical discharge, and return to the ground state by emitting yellow light of wavelength 589nm. Usually the excitation of valence electrons (such as 3s for sodium) involves energies corresponding to photons of visible or ultraviolet light. The excitation of core electrons is possible, but requires much higher energies generally corresponding to x-ray photons. This would be the case for example to excite a 2p electron to the 3s level and form the excited 1s22s22p53s2 configuration. The remainder of this article deals only with the ground-state configuration, often referred to as "the" configuration of an atom or molecule.

History
Niels Bohr was the first to propose (1923) that the periodicity in the properties of the elements might be explained by the electronic structure of the atom.[7] His proposals were based on the then current Bohr model of the atom, in which the electron shells were orbits at a fixed distance from the nucleus. Bohr's original configurations would seem strange to a present-day chemist: sulfur was given as 2.4.4.6 instead of 1s22s22p63s23p4 (2.8.6). The following year, E.C.Stoner incorporated Sommerfeld's third quantum number into the description of electron shells, and correctly predicted the shell structure of sulfur to be 2.8.6.[8] However neither Bohr's system nor Stoner's could correctly describe the changes in atomic spectra in a magnetic field (the Zeeman effect). Bohr was well aware of this shortcoming (and others), and had written to his friend Wolfgang Pauli to ask for his help in saving quantum theory (the system now known as "old quantum theory"). Pauli realized that the Zeeman effect must be due only to the outermost electrons of the atom, and was able to reproduce Stoner's shell structure, but with the correct structure of subshells, by his inclusion of a fourth quantum number and his exclusion principle (1925):[9] It should be forbidden for more than one electron with the same value of the main quantum number n to have the same value for the other three quantum numbers k [l], j [ml] and m [ms]. The Schrdinger equation, published in 1926, gave three of the four quantum numbers as a direct consequence of its solution for the hydrogen atom:[2] this solution yields the atomic orbitals which are shown today in textbooks of chemistry (and above). The examination of atomic spectra allowed the electron configurations of atoms to be determined experimentally, and led to an empirical rule (known as Madelung's rule (1936),[10] see below) for the order in which atomic orbitals are filled with electrons.

Electron configuration

245

Aufbau principle and Madelung rule


The Aufbau principle (from the German Aufbau, "building up, construction") was an important part of Bohr's original concept of electron configuration. It may be stated as:[11] a maximum of two electrons are put into orbitals in the order of increasing orbital energy: the lowest-energy orbitals are filled before electrons are placed in higher-energy orbitals. The principle works very well (for the ground states of the atoms) for the first 18elements, then decreasingly well for the following 100elements. The modern form of the Aufbau principle describes an order of orbital energies given by Madelung's rule (or Klechkowski's rule). This rule was first stated by Charles Janet in 1929, rediscovered by Erwin Madelung in 1936,[10] and later given a theoretical justification by V.M. Klechkowski[12]

The approximate order of filling of atomic orbitals, following the arrows from 1s to 7p. (After 7p the order includes orbitals outside the range of the diagram, starting with 8s.)

1. Orbitals are filled in the order of increasing n+l; 2. Where two orbitals have the same value of n+l, they are filled in order of increasing n. This gives the following order for filling the orbitals: 1s, 2s, 2p, 3s, 3p, 4s, 3d, 4p, 5s, 4d, 5p, 6s, 4f, 5d, 6p, 7s, 5f, 6d, 7p, (8s, 5g, 6f, 7d, 8p, and 9s) In this list the orbitals in parentheses are not occupied in the ground state of the heaviest atom now known (Uuo, Z = 118). The Aufbau principle can be applied, in a modified form, to the protons and neutrons in the atomic nucleus, as in the shell model of nuclear physics and nuclear chemistry.

Periodic table
The form of the periodic table is closely related to the electron configuration of the atoms of the elements. For example, all the elements of group 2 have an electron configuration of [E]ns2 (where [E] is an inert gas configuration), and have notable similarities in their chemical properties. In general, the periodicity of the periodic table in terms of periodic table blocks is clearly due to the number of electrons (2, 6, 10, 14...) needed to fill s, p, d, and f subshells. The outermost electron shell is often referred to as the "valence shell" and (to a Electron configuration table first approximation) determines the chemical properties. It should be remembered that the similarities in the chemical properties were remarked more than a century before the idea of electron configuration,[13] It is not clear how far Madelung's rule explains (rather than simply describes) the periodic table,[14] although some properties (such as the common +2 oxidation state in the first row of the transition metals) would obviously be different with a different order of orbital filling.

Electron configuration

246

Shortcomings of the Aufbau principle


The Aufbau principle rests on a fundamental postulate that the order of orbital energies is fixed, both for a given element and between different elements: neither of these is true (although they are approximately true enough for the principle to be useful). It considers atomic orbitals as "boxes" of fixed energy into which can be placed two electrons and no more. However the energy of an electron "in" an atomic orbital depends on the energies of all the other electrons of the atom (or ion, or molecule, etc.). There are no "one-electron solutions" for systems of more than one electron, only a set of many-electron solutions which cannot be calculated exactly[15] (although there are mathematical approximations available, such as the HartreeFock method). The fact that the Aufbau principle is based on an approximation can be seen from the fact that there is an almost-fixed filling order at all, that, within a given shell, the s-orbital is always filled before the p-orbitals. In a hydrogen-like atom, which only has one electron, the s-orbital and the p-orbitals of the same shell have exactly the same energy, to a very good approximation in the absence of external electromagnetic fields. (However, in a real hydrogen atom, the energy levels are slightly split by the magnetic field of the nucleus, and by the quantum electrodynamic effects of the Lamb shift).

Ionization of the transition metals


The naive application of the Aufbau principle leads to a well-known paradox (or apparent paradox) in the basic chemistry of the transition metals. Potassium and calcium appear in the periodic table before the transition metals, and have electron configurations [Ar]4s1 and [Ar]4s2 respectively, i.e. the 4s-orbital is filled before the 3d-orbital. This is in line with Madelung's rule, as the 4s-orbital has n+l = 4 (n= 4, l= 0) while the 3d-orbital has n+l = 5 (n= 3, l= 2). However, chromium and copper have electron configurations [Ar]3d54s1 and [Ar]3d104s1 respectively, i.e. one electron has passed from the 4s-orbital to a 3d-orbital to generate a half-filled or filled subshell. In this case, the usual explanation is that "half-filled or completely filled subshells are particularly stable arrangements of electrons". The apparent paradox arises when electrons are removed from the transition metal atoms to form ions. The first electrons to be ionized come not from the 3d-orbital, as one would expect if it were "higher in energy", but from the 4s-orbital. The same is true when chemical compounds are formed. Chromium hexacarbonyl can be described as a chromium atom (not ion, it is in the oxidation state0) surrounded by six carbon monoxide ligands: it is diamagnetic, and the electron configuration of the central chromium atom is described as 3d6, i.e. the electron which was in the 4s-orbital in the free atom has passed into a 3d-orbital on forming the compound. This interchange of electrons between 4s and 3d is universal among the first series of the transition metals.[16] The phenomenon is only paradoxical if it is assumed that the energies of atomic orbitals are fixed and unaffected by the presence of electrons in other orbitals. If that were the case, the 3d-orbital would have the same energy as the 3p-orbital, as it does in hydrogen, yet it clearly doesn't. There is no special reason why the Fe2+ ion should have the same electron configuration as the chromium atom, given that iron has two more protons in its nucleus than chromium and that the chemistry of the two species is very different. When care is taken to compare "like with like", the paradox disappears.[17]

Other exceptions to Madelung's rule


There are several more exceptions to Madelung's rule among the heavier elements, and it is more and more difficult to resort to simple explanations such as the stability of half-filled subshells. It is possible to predict most of the exceptions by HartreeFock calculations,[18] which are an approximate method for taking account of the effect of the other electrons on orbital energies. For the heavier elements, it is also necessary to take account of the effects of Special Relativity on the energies of the atomic orbitals, as the inner-shell electrons are moving at speeds approaching the speed of light. In general, these relativistic effects[19] tend to decrease the energy of the s-orbitals in relation to the other atomic orbitals.[20]

Electron configuration

247

Electron shells filled in violation of Madelung's rule[21] (red)


Period 4 Element Z Electron Element Configuration Period 5 Z Electron Element Configuration Lanthanum Cerium Period 6 Z Electron Element Configuration Actinium Thorium Period 7 Z Electron Configuration [Rn] 7s2 6d1 [Rn] 7s2 6d2

57 [Xe] 6s2 5d1 58 [Xe] 6s2 4f1 5d1

89 90

Praseodymium 59 [Xe] 6s2 4f3

Protactinium

91

[Rn] 7s2 5f2 6d1 [Rn] 7s2 5f3 6d1 [Rn] 7s2 5f4 6d1 [Rn] 7s2 5f6 [Rn] 7s2 5f7 [Rn] 7s2 5f7 6d1 [Rn] 7s2 5f9

Neodymium

60 [Xe] 6s2 4f4

Uranium

92

Promethium

61 [Xe] 6s2 4f5

Neptunium

93

Samarium Europium Gadolinium

62 [Xe] 6s2 4f6 63 [Xe] 6s2 4f7 64 [Xe] 6s2 4f7 5d1 65 [Xe] 6s2 4f9 71 [Xe] 6s2 4f14 5d1 72 [Xe] 6s2 4f14 5d2 73 [Xe] 6s2 4f14 5d3 74 [Xe] 6s2 4f14 5d4 75 [Xe] 6s2 4f14 5d5 76 [Xe] 6s2 4f14 5d6 77 [Xe] 6s2 4f14 5d7 78 [Xe] 6s1 4f14 5d9

Plutonium Americium Curium

94 95 96

Terbium Scandium 21 [Ar] 4s2 3d1 Yttrium 39 [Kr] 5s2 4d1 Lutetium

Berkelium Lawrencium

97

103 [Rn] 7s2 5f14 7p1

Titanium

22 [Ar] 4s2 3d2

Zirconium

40 [Kr] 5s2 4d2

Hafnium

Rutherfordium 104 [Rn] 7s2 5f14 6d2

Vanadium

23 [Ar] 4s2 3d3

Niobium

41 [Kr] 5s1 4d4

Tantalum

Chromium 24 [Ar] 4s1 3d5

Molybdenum 42 [Kr] 5s1 4d5

Tungsten

Manganese 25 [Ar] 4s2 3d5

Technetium

43 [Kr] 5s2 4d5

Rhenium

Iron

26 [Ar] 4s2 3d6

Ruthenium

44 [Kr] 5s1 4d7

Osmium

Cobalt

27 [Ar] 4s2 3d7

Rhodium

45 [Kr] 5s1 4d8

Iridium

Nickel

28 [Ar] 4s2 3d8 or [Ar] 4s1 3d9 [22] (disputed) 29 [Ar] 4s1 3d10

Palladium

46 [Kr] 4d10

Platinum

Copper

Silver

47 [Kr] 5s1 4d10

Gold

79 [Xe] 6s1 4f14 5d10 80 [Xe] 6s2 4f14 5d10

Zinc

30 [Ar] 4s2 3d10

Cadmium

48 [Kr] 5s2 4d10

Mercury

The electron-shell configuration of elements beyond rutherfordium is not yet known.

Electron configuration

248

Electron configuration in molecules


In molecules, the situation becomes more complex, as each molecule has a different orbital structure. The molecular orbitals are labelled according to their symmetry,[23] rather than the atomic orbital labels used for atoms and monoatomic ions: hence, the electron configuration of the dioxygen molecule, O2, is 1g21u22g22u21u43g21g2.[1] The term 1g2 represents the two electrons in the two degenerate *-orbitals (antibonding). From Hund's rules, these electrons have parallel spins in the ground state, and so dioxygen has a net magnetic moment (it is paramagnetic). The explanation of the paramagnetism of dioxygen was a major success for molecular orbital theory.

Electron configuration in solids


In a solid, the electron states become very numerous. They cease to be discrete, and effectively blend into continuous ranges of possible states (an electron band). The notion of electron configuration ceases to be relevant, and yields to band theory.

Applications
The most widespread application of electron configurations is in the rationalization of chemical properties, in both inorganic and organic chemistry. In effect, electron configurations, along with some simplified form of molecular orbital theory, have become the modern equivalent of the valence concept, describing the number and type of chemical bonds that an atom can be expected to form. This approach is taken further in computational chemistry, which typically attempts to make quantitative estimates of chemical properties. For many years, most such calculations relied upon the "linear combination of atomic orbitals" (LCAO) approximation, using an ever larger and more complex basis set of atomic orbitals as the starting point. The last step in such a calculation is the assignment of electrons among the molecular orbitals according to the Aufbau principle. Not all methods in calculational chemistry rely on electron configuration: density functional theory (DFT) is an important example of a method which discards the model. A fundamental application of electron configurations is in the interpretation of atomic spectra. In this case, it is necessary to convert the electron configuration into one or more term symbols, which describe the different energy levels available to an atom. Term symbols can be calculated for any electron configuration, not just the ground-state configuration listed in tables, although not all the energy levels are observed in practice. It is through the analysis of atomic spectra that the ground-state electron configurations of the elements were experimentally determined.

Notes
[1] Nic, M.; Jirat, J.; Kosata, B., eds. (2006). "configuration (electronic)" (http:/ / goldbook. iupac. org/ C01248. html). IUPAC Compendium of Chemical Terminology (Online ed.). doi:10.1351/goldbook.C01248. ISBN0-9678550-9-8. . [2] In formal terms, the quantum numbers n, l and ml arise from the fact that the solutions to the time-independent Schrdinger equation for hydrogen-like atoms are based on spherical harmonics. [3] Nic, M.; Jirat, J.; Kosata, B., eds. (2006). "Pauli exclusion principle" (http:/ / goldbook. iupac. org/ PT07089. html). IUPAC Compendium of Chemical Terminology (Online ed.). doi:10.1351/goldbook.PT07089. ISBN0-9678550-9-8. . [4] The full form of the configuration notation is a mathematical product, so 3p3 indicates that it is the cube of the 3p function which enters into the product (even if it is not normal to pronounce it in that way). [5] Weisstein, Eric W. (2007). "Electron Orbital" (http:/ / scienceworld. wolfram. com/ physics/ ElectronOrbital. html). wolfram. . [6] Ebbing, Darrell D.; Gammon, Steven D. (2007-01-12). General Chemistry (http:/ / books. google. com/ ?id=_vRm5tiUJcsC& pg=PA284& lpg=PA284& dq=choice+ of+ letters+ s+ p+ d+ orbitals+ diffuse#v=onepage& q& f=false). p.284. ISBN978-0-618-73879-3. . [7] Bohr, Niels (1923). "ber die Anwendung der Quantumtheorie auf den Atombau. I". Zeitschrift fr Physik 13: 117. Bibcode1923ZPhy...13..117B. doi:10.1007/BF01328209. [8] Stoner, E.C. (1924). "The distribution of electrons among atomic levels". Philosophical Magazine (6th Ser.) 48 (286): 71936. doi:10.1080/14786442408634535.

Electron configuration
[9] Pauli, Wolfgang (1925). "ber den Einfluss der Geschwindigkeitsabhndigkeit der elektronmasse auf den Zeemaneffekt". Zeitschrift fr Physik 31: 373. Bibcode1925ZPhy...31..373P. doi:10.1007/BF02980592. English translation from Scerri, Eric R. (1991). "The Electron Configuration Model, Quantum Mechanics and Reduction" (http:/ / www. chem. ucla. edu/ dept/ Faculty/ scerri/ pdf/ BJPS. pdf). Br. J. Phil. Sci. 42 (3): 30925. doi:10.1093/bjps/42.3.309. . [10] Madelung, Erwin (1936). Mathematische Hilfsmittel des Physikers. Berlin: Springer. [11] Nic, M.; Jirat, J.; Kosata, B., eds. (2006). "aufbau principle" (http:/ / goldbook. iupac. org/ AT06996. html). IUPAC Compendium of Chemical Terminology (Online ed.). doi:10.1351/goldbook.AT06996. ISBN0-9678550-9-8. . [12] Wong, D. Pan (1979). "Theoretical justification of Madelung's rule" (http:/ / jchemed. chem. wisc. edu/ Journal/ Issues/ 1979/ Nov/ jceSubscriber/ JCE1979p0714. pdf). Journal of Chemical Education 56 (11): 71418. Bibcode1979JChEd..56..714W. doi:10.1021/ed056p714. . [13] The similarities in chemical properties and the numerical relationship between the atomic weights of calcium, strontium and barium was first noted by Johann Wolfgang Dbereiner in 1817. [14] Scerri, Eric R. (1998). "How Good Is the Quantum Mechanical Explanation of the Periodic System?" (http:/ / www. chem. ucla. edu/ dept/ Faculty/ scerri/ pdf/ How_Good_is. pdf). Journal of Chemical Education 75 (11): 138485. Bibcode1998JChEd..75.1384S. doi:10.1021/ed075p1384. . Ostrovsky, V.N. (2005). "On Recent Discussion Concerning Quantum Justification of the Periodic Table of the Elements" (http:/ / www. springerlink. com/ content/ p2rqg32684034736/ fulltext. pdf). Foundations of Chemistry 7 (3): 23539. doi:10.1007/s10698-005-2141-y. . Abstract (http:/ / www. springerlink. com/ content/ p2rqg32684034736/ fulltext. pdf?page=1). [15] Electrons are identical particles, a fact which is sometimes referred to as "indistinguishability of electrons". A one-electron solution to a many-electron system would imply that the electrons could be distinguished from one another, and there is strong experimental evidence that they can't be. The exact solution of a many-electron system is a n-body problem with n 3 (the nucleus counts as one of the "bodies"): such problems have evaded analytical solution since at least the time of Euler. [16] There are some cases in the second and third series where the electron remains in an s-orbital. [17] Melrose, Melvyn P.; Scerri, Eric R. (1996). "Why the 4s Orbital is Occupied before the 3d" (http:/ / jchemed. chem. wisc. edu/ Journal/ Issues/ 1996/ Jun/ jceSubscriber/ JCE1996p0498. pdf). Journal of Chemical Education 73 (6): 498503. Bibcode1996JChEd..73..498M. doi:10.1021/ed073p498. . Abstract (http:/ / adsabs. harvard. edu/ abs/ 1996JChEd. . 73. . 498M). [18] Meek, Terry L.; Allen, Leland C. (2002). "Configuration irregularities: deviations from the Madelung rule and inversion of orbital energy levels" (http:/ / www. sciencedirect. com/ science?_ob=ArticleURL& _udi=B6TFN-46G4S5S-1& _user=961305& _rdoc=1& _fmt=& _orig=search& _sort=d& view=c& _acct=C000049425& _version=1& _urlVersion=0& _userid=961305& md5=cef78ae6aced8ded250c6931a0842063). Chem. Phys. Lett. 362 (56): 36264. Bibcode2002CPL...362..362M. doi:10.1016/S0009-2614(02)00919-3. . [19] Nic, M.; Jirat, J.; Kosata, B., eds. (2006). "relativistic effects" (http:/ / goldbook. iupac. org/ RT07093. html). IUPAC Compendium of Chemical Terminology (Online ed.). doi:10.1351/goldbook.RT07093. ISBN0-9678550-9-8. . [20] Pyykk, Pekka (1988). "Relativistic effects in structural chemistry". Chem. Rev. 88 (3): 56394. doi:10.1021/cr00085a006. [21] G.L. Miessler and D.A. Tarr, "Inorganic Chemistry" (2nd ed., Prentice-Hall 1999) p.38 [22] Scerri, Eric R. (2007). The periodic table: its story and its significance (http:/ / books. google. com/ ?id=SNRdGWCGt1UC& pg=PA239). Oxford University Press. pp.239240. ISBN0-19-530573-6. . [23] The labels are written in lowercase to indicate that the correspond to one-electron functions. They are numbered consecutively for each symmetry type (irreducible representation in the character table of the point group for the molecule), starting from the orbital of lowest energy for that type.

249

References
Jolly, William L. (1991). Modern Inorganic Chemistry (2nd ed.). New York: McGraw-Hill. pp.123. ISBN0-07-112651-1. Scerri, Eric (2007). The Periodic System, Its Story and Its Significance. New York: Oxford University Press. ISBN0-19-530573-6.

External links
What does an atom look like? Configuration in 3D (http://www.hydrogenlab.de/elektronium/HTML/ einleitung_hauptseite_uk.html)

Inverse magnetostrictive effect

250

Inverse magnetostrictive effect


The inverse magnetostrictive effect (also known as Villari effect) is the name given to the change of the magnetic susceptibility of a material when subjected to a mechanical stress.

Explanation
Whereas magnetostriction characterizes the shape change of a ferromagnetic material during magnetization, the inverse magnetostrictive effect characterizes the change of domain magnetization when a stress is applied to a material. This magnetostriction can be positive (magnetization increased by tension) like in pure iron, or negative (magnetization decreased by tension) like in nickel. In the case of a single stress applied on a single magnetic domain, the magnetic strain energy density can be expressed as [1] :

where

is the magnetostrictive expansion at saturation, and and

the angle between the saturation magnetization and =

the stressed direction. When

are both positive (like in iron under tension), the energy is minimum for

0, i.e. when tension is aligned with the saturation magnetization. Consequently, the magnetization is increased by tension. In fact, magnetostriction is more complex and depends on the direction of the crystal axes. In iron, the [100] axes are the directions of easy magnetization, while there is little magnetization along the [111] directions (unless the magnetization becomes close to the saturation magnetization, leading to the change of the domain orientation from [111] to [100]). This magnetic anisotropy pushed authors to define two independent longitudinal magnetostrictions and . In cubic materials, the magnetostriction along any axis can be defined by a known linear combination of these two constants. For instance, the elongation along [110] is a linear combination of and . Under assumptions of isotropic magnetostriction (i.e. domain magnetization is the same in any crystallographic directions), then and the linear dependence between the elastic energy and the stress is conserved, domain magnetization, and , , . Here, , and are the direction cosines of the

those of the bond directions, towards the crystallographic directions.

References
[1] Bozorth, R. (1951). Ferromagnetism. Van Nostrand.

Exchange bias

251

Exchange bias
Exchange bias or exchange anisotropy occurs in bilayers (or multilayers) of magnetic materials where the hard magnetization behavior of an antiferromagnetic thin film causes a shift in the soft magnetization curve of a ferromagnetic film. The exchange bias phenomenon is of tremendous utility in magnetic recording, where it is used to pin the state of the readback heads of hard disk drives at exactly their point of maximum sensitivity; hence the term "bias."

Fundamental science
The essential physics underlying the phenomenon is the exchange interaction between the antiferromagnet and ferromagnet at their interface. Since antiferromagnets have a small or no net magnetization, their spin orientation is only weakly influenced by an externally applied magnetic field. A soft ferromagnetic film which is strongly Easy-axis magnetization curves of a) a soft exchange-coupled to the antiferromagnet will have its interfacial spins ferromagnetic film; b) an antiferromagnetic film and c) an exchange-biased bilayer consisting of a pinned. Reversal of the ferromagnet's moment will have an added ferromagnet and an antiferromagnet. The energetic cost corresponding to the energy necessary to create a Nel susceptibility (slope) of the antiferromagnetic's domain wall within the ferromagnetic film. The added energy term magnetization curve is exaggerated for clarity. implies a shift in the switching field of the ferromagnet. Thus the magnetization curve of an exchange-biased ferromagnetic film looks like that of the normal ferromagnet except that is shifted away from the H=0 axis by an amount Hb. In most well-studied ferromagnet/antiferromagnet bilayers, the Curie temperature of the ferromagnet is larger than the Nel temperature TN of the antiferromagnet. This inequality means that the direction of the exchange bias can be set by cooling through TN in the presence of an applied magnetic field. The moment of the magnetically ordered ferromagnet will apply an effective field to the antiferromagnet as it orders, breaking the symmetry and influencing the formation of domains. Exchange anisotropy has long been poorly understood due to the difficulty of studying the dynamics of domain walls in thin antiferromagnetic films. A naive approach to the problem would suggest the following expression for energy per unit area:

where n is the number of interfacial spins interactions per unit area, Jex is the exchange constant at the interface, S refers to the spin vector, M refers to the magnetization, t refers to film thickness and H is the external field. The subscript F describes the properties of the ferromagnet and AF to the antiferromagnet. The expression omits magnetocrystalline anisotropy, which is unaffected by the presence of the antiferromagnet. At the switching field of the ferromagnet, the pinning energy represented by the first term and the Zeeman dipole coupling represented by the second term will exactly balance. The equation then predicts that the exchange bias shift Hb will be given by the expression

Many experimental findings regarding the exchange bias contradict this simple model. For example, the magnitude of measured Hb values is typically 100 times less than that predicted by the equation for reasonable values of the parameters. The amount of hysteresis shift Hb is not correlated with the density n of uncompensated spins in the plane of the antiferromagnet that appears at the interface. In addition, the exchange bias effect tends to be smaller in epitaxial bilayers than in polycrystalline ones, suggesting an important role for defects. In recent years progress in

Exchange bias fundamental understanding has been made via synchrotron radiation based element-specific magnetic linear dichroism experiments that can image antiferromagnetic domains and frequency-dependent magnetic susceptibility measurements that can probe the dynamics. Experiments on the Fe/FeF2 and Fe/MnF2 model systems have been particularly fruitful.

252

Technological impact
Exchange bias was initially used to stabilize the magnetization of soft ferromagnetic layers in readback heads based on the anisotropic magnetoresistance (AMR) effect. Without the stabilization, the magnetic domain state of the head could be unpredictable, leading to reliability problems. Currently exchange bias is used to pin the harder reference layer in spin valve readback heads and MRAM memory circuits that utilize the giant magnetoresistance or magnetic tunneling effect. Similarly the most advanced disk media are antiferromagnetically coupled, making use of interfacial exchange to effectively increase the stability of small magnetic particles whose behavior would otherwise be superparamagnetic. Desirable properties for an exchange bias material include a high Nel temperature, a large magnetocrystalline anisotropy and good chemical and structural compatibility with NiFe and Co, the most important ferromagnetic films. The most technologically significant exchange bias materials have been the rocksalt-structure antiferromagnetic oxides like NiO, CoO and their alloys and the rocksalt-structure intermetallics like FeMn, NiMn, IrMn and their alloys.

History
Exchange anisotropy was discovered by Meiklejohn and Bean of General Electric in 1956. The first commercial device to employ the exchange bias was IBM's anisotropic magnetoresistance (AMR) disk drive recording head, which was based on a design by Hunt in the 1970s but which didn't fully displace the inductive readback head until the early 1990s. By the mid-1990s, the spin valve head using an exchange-bias layer was well on its way to displacing the AMR head.

Article Sources and Contributors

253

Article Sources and Contributors


Magnetism Source: http://en.wikipedia.org/w/index.php?oldid=489348793 Contributors: 16@r, 2over0, 4lex, 5 albert square, A Mom, A8UDI, AL2TB, Abce2, Adawg117, Addshore, Ahoerstemeier, Akendall, Akriasas, Aksi great, Al.locke, Alansohn, Ale jrb, Alex-engraver, Aliwikii, Allstarecho, American Eagle, Amog, Andonic, Andy Dingley, Anonymous Dissident, Antandrus, Apocalypse2009, Bassbonerocks, Bcfootball, Bchaplucian, Beland, Benbest, Bensaccount, Blizzarex, Bob sagget jr., Bobo192, Bongwarrior, Boredzo, Brews ohare, Brianga, Brockert, Bryan Derksen, C d h, CDN99, CYD, Caknuck, CalebNoble, Caltas, Calvin 1998, Can't sleep, clown will eat me, Capricorn42, Capybara123, Casull, Catgut, Caturdayz, Celarnor, Chaiken, Changer-guy, Charivari, Charles Matthews, Chaser, Chetvorno, Chrislk02, ChucksGay123456789, Chzz, Clark89, ClarkLewis, Closedmouth, Cmandouble3, Cmichael, Cntras, CommonsDelinker, Conversion script, Craig Pemberton, Crawlbeforeiwalk, Cryptoid, Ctachme, D6, DJIndica, DMacks, DV8 2XL, Dalegudmunsen, Dan Austin, Dan kelley90, DanMatan, Daniel.Cardenas, Davdclehn, David.Mestel, Davidhorman, Davidkazuhiro, Dbtfz, Delirium, DerHexer, Dfrg.msc, Dina, Discospinster, Donarreiskoffer, Doodoobutter, Doulos Christos, Dr. Sunglasses, Drostie, Drrngrvy, Dubbin, Duncan.france, Duncharris, Dugosz, Dugosz, E23, E2eamon, ESkog, EWikist, Ed Poor, Edderso, Edinborgarstefan, Edison, Edward, Edward Z. Yang, Eeekster, Egil, Egmontaz, El aprendelenguas, ElTyrant, Electrodynamicist, ElectronicsEnthusiast, Elipongo, Ellywa, Enormousdude, Epbr123, EquinoX, Erniesaurus, Eubulides, FJPB, Fang Aili, Feline1, Femto, Ferix, Firozmusthafa, Fizped, Flubber88, Fluidchameleon, Flutterman, Fredrik, Froid, Fuhghettaboutit, Gail, Gaius Cornelius, Gangasudhan, Gene Nygaard, Geologyguy, George2001hi, GhostPirate, Giftlite, Gioto, Glacialfox, Glenn, Gnowor, Gogo Dodo, GoldenTorc, GraemeL, Grapetonix, Griffin5, Gscshoyru, Guanaco, Gurch, Gzuckier, Haham hanuka, Hairy Dude, Halpaugh, Harold f, Headbomb, Hellbus, HereToHelp, Heron, HexaChord, Hmains, Hobartimus, Hongooi, Hu, Hv, Hydrogen Iodide, II MusLiM HyBRiD II, IMNOTARETARDATALL, IW.HG, Icairns, IceUnshattered, Immunize, IndulgentReader, Insanity Incarnate, Inthepink, InverseHypercube, Ipigott, Iridescent, Isis, Ixfd64, J. Spencer, J.delanoy, JA.Davidson, JCPH, JDspeeder1, JForget, JabberWok, JacobTrue, Jaeger5432, Jagged 85, Jakhai1000, Jamesontai, Jauhienij, Jcwf, Jeffhoy, Jeremy Visser, Jessemv, Jfdwolff, Jim.belk, Jmabel, Jmundo, Joanjoc, Jocelyne Heys-Gerard, John Aplessed, JorisvS, JulianB12, Junglecat, Jusdafax, JustAddPeter, Karch, Karcih, Karl-Henner, Karol Langner, Karteek987, Kazvorpal, Keegan, Keilana, Kerotan, Khyranleander, King of Hearts, Kingpin13, Kmarinas86, Knutsi, Koavf, Kostisl, Ktsquare, KyleCardoza, Lahiru k, Lankiveil, Leif27, Light current, Lightmouse, Lir, Llywrch, Lone Skeptic, Lordyou, LorenzoB, M412k, MELISASIMPSONS, MER-C, MZMcBride, Mac, Magister Mathematicae, MarcoAurelio, Martarius, Masonm95, Masonprof, Maxellus, Maxrokatanski, Mdanziger, Meggar, Melchoir, Melongrower, Mentifisto, Mephistophelian, Metacomet, Michael Hardy, Mike Dill, Mike by, Mistercow, Mjspe1, Mkch, Mobile Snail, Modemac, Mp50967, Mr Stephen, Mr. Trustegious, Mspraveen, Mufka, MusicMaker5376, Mxn, Mygerardromance, NHRHS2010, Nabla, Natalie Erin, Nauticashades, NawlinWiki, NewEnglandYankee, Nick Number, Nicktfx, Nihiltres, Ninjackster, Nmgrad, Nmnogueira, Noor Qasmieh, Nsaa, Octahedron80, Oliviosu, Omicronpersei8, Opelio, Otolemur crassicaudatus, Ottawa4ever, Ownedestroy, Ozone77, Ozuma, Paddy-B-Jr, Pak21, Paolo.dL, ParkerHiggins, Party, Pedant17, Peterlin, Petr Kopa, Petrb, Pgk, Pharaoh of the Wizards, Philip Trueman, Physchim62, Piano non troppo, Pigman, Pilotguy, Pinestone, Piotrus, Pjbcool103, Pmbeck, Pointillist, PoolDoc, Postscript07, PrincessofLlyr, ProDigit, Proficient, Proofreader77, Puchiko, Pudgy78685, PwncakesN bacon, Qui-Gon Jinn, Quickbeam, Quintote, RDR, RG2, Radon210, Rafonseca, Raminmahpour, RaulRavndra, Reconsider the static, Reddi, Redfarmer, Renesis, RexNL, Rifleman 82, Rising*From*Ashes, Roberdor, Robma, RockMagnetist, Rolinator, Ron B. Thomson, Rossf18, Royboycrashfan, Rpb01r, Rrburke, Russell4, Ryalisaivamsi, Ryne1, S colligan, S ortiz, SGMD1, Sadi Carnot, Saideepak.budaraju, Sammyk214, Samsee, Sandstein, Satanael, Sbyrnes321, Schultz.Ryan, Science5, Scientizzle, Scott3, Shadowjams, Shotwell, Shrigley, Shuipzv3, Sietse Snel, Silent78, Sinewalker, Sionus, Skew-t, Skizzik, Slazenger, Slgrandson, Slon02, Sluzzelin, Smack, Smilesfozwood, Snowmanradio, Snowolf, Snoyes, Some jerk on the Internet, Someguy1221, Spangineer, Special-T, Spencer, Spinningspark, Splartmaggot, Srikar33, Ssilvers, St. Hubert, Stanlste, Stephenb, SteveBaker, Stevertigo, Stokerm, Stone, Strait, Stuart07, Suffusion of Yellow, Sunray, Syrthiss, THEN WHO WAS PHONE?, TStein, Techguru, Techman224, Technobebop, TehBrandon, Template namespace initialisation script, Tempodivalse, Texture, Tgv8925, That Guy, From That Show!, The Rambling Man, The Thing That Should Not Be, The Troll lolololololololol, TheBendster, TheFronze05, TheMolecularMan, Thekillerpenguin, Theresa knott, Thingg, This user has left wikipedia, Thisisborin9, Tiddly Tom, Tide rolls, Tim Shuba, Tim Starling, Tobby72, Tobias Bergemann, Tommy2010, Trojancowboy, Tyler Matthews, Ulric1313, Unyoyega, Usp, Uvainio, VASANTH S.N., Vadim Makarov, Vanilluv30, Vanished user 39948282, VanishedUser314159, Versus22, Viriditas, Vrenator, Vsmith, Vssun, WOSlinker, Warut, Waveguy, Wayne Slam, Wenli, Where, Whereizben, WikHead, Wiki alf, Wikier.ko, William S. Saturn, Willking1979, Wolfkeeper, Woudloper, Wtmitchell, Xenonice, Xxanthippe, Yamamoto Ichiro, Yammer68, Yevgeny Kats, Ykral, Yoduh2007, Yoyo2222, Yuwangswisscom, Z.E.R.O., Zoragotcha, Zundark, Zvika, , 1212 anonymous edits Magnetic field Source: http://en.wikipedia.org/w/index.php?oldid=489809323 Contributors: 1994bhaskar, 1howardsr1, 213.253.39.xxx, 23790AD, 2D, 2over0, 4twenty42o, @pple, Abductive, Af648, Ahoerstemeier, Aitias, Aka042, Alansohn, Alex Klotz, Alfred Centauri, Alphachimp, Ambros-aba, Ankid, Anna512, Anterior1, Antixt, Arch dude, Armius, Arthena, Ascidian, Ashill, Aulis Eskola, B21O303V3941W42371, BSTR, Bachrach44, Barneca, Bart133, Bekus, Beland, BenFrantzDale, Bender235, Bishoppowell, Black Shadow, Bobyorox, Bookandcoffee, Brews ohare, BrianWilloughby, Brichcja, Brigman, Bryan Derksen, Buster79, C14, CUSENZA Mario, Calvin 1998, Cantiorix, Capricorn42, Captain Yankee, CaptinJohn, Catslash, Celebere, Cessator, CharlesChandler, Chetvorno, Chris the speller, Chrislk02, Chrsschm, Citeseer, Clicketyclack, CliffC, Clw, Cocytus, Coldwarrier, Cometstyles, Complexica, CosineKitty, Creidieki, Curps, Cxz111, CyrilB, D6, DARTH SIDIOUS 2, DJIndica, DMahalko, Da Joe, Dadude3320, Daf, Damo0078, Daniel.Cardenas, DeMk9D76, Deadlyops, Defender of torch, Delirium, DemonThing, Dennis Brown, DerHexer, DesertAngel, Dgmyer, Dgroseth, Diannaa, Dicklyon, Dino, Direvus, Discospinster, Djr32, Dmn, Doc aberdeen, DocWatson42, DomenicDenicola, DoubleBlue, Download, Doyley, Dr. Seaweed, DrBob, Dreadengineer, Drrngrvy, Dynaflow, Edivorce, Egmontaz, Ekkert, El C, El estremeu, Enchanter, Enormousdude, Ethan, Ettrig, Eudoxie, Evgeny, Excirial, Explodinglam, FDT, FelisLeo, Femto, Filemon, Filippopoulos, Floorsheim, Fongs, Fpahl, Freelance Physicist, Fresheneesz, From-cary, FrstFrs, Fyyer, FyzixFighter, F, G-W, GDonato, GRB, Gaius Cornelius, Gary King, Gatoatigrado, Geek1337, Gene Nygaard, Gfoley4, Giftlite, Giggy12345, Gilliam, Gingavitus777, Gits (Neo), Glenn, Glmory, Goudzovski, Greenpowered, Grstain, Gryllida, Gkhan, H0dges, H2g2bob, HEL, Harry, Headbomb, Helix84, Hellbus, HenryLi, Herbee, Heron, HexaChord, Hommadi2001, Hongooi, Hqb, Hunter360x, Hurricane Angel, Hydrogen Iodide, ICE77, IVAN3MAN, Iantresman, Icairns, Icep, Imafjbks, Imrankhan85, Incompetence, Inquisitus, Intangir, InverseHypercube, Iridescent, Isdarts222, Ixfd64, J.delanoy, JD554, JForget, JJ Harrison, JaGa, JabberWok, Jackelfive, JackyR, Jacobymathews99, Jaganath, Jagun, Jakebarrington, Janolaf30, JasonSaulG, Jauhienij, Javalenok, Jaxl, JerrySteal, Jfx319, Jim1138, JoanneB, John of Reading, John254, JohnBlackburne, Jojalozzo, Justanyone, Jvansanten, KJS77, Kafka Liz, KasugaHuang, Katalaveno, Katieh5584, Kenshin9554, Kesac, Khashishi, Kimse, Kingpin13, Kku, Kmarinas86, Kri, Kurt.hewett, Kurzon, Laurascudder, LeCire, LeilaniLad, Lenko, Lichen from Hell, Light current, LilHelpa, Lindberg G Williams Jr, Locriani, Lookang, Lseixas, Luke490, Luna Santin, M C Y 1008, MC10, MER-C, MacedonianBoy, Magnetic7, Magog the Ogre, Manscher, MarcoLittel, Marmzok, MarsRover, Maschen, Matdrodes, Materialscientist, Mattbr, Maxhutch, Mcmonkeyburger, Mebden, Melchoir, Mentifisto, Merseyless, Metacomet, Mh liv01, Michael Hardy, Michi zh, Mihail Vasiliev, Mikiemike, Mjpieters, Mni9791, Modulatum, Momo san, Mossd, Msiddalingaiah, Mstyne, Myasuda, Mygerardromance, Nabla, Nageh, Nakon, NatureA16, Ndhuang, Neko-chan, Netheril96, NewEnglandYankee, Nicolharper, Nielchiano, Nigilan, Nmnogueira, Noommos, Notinasnaid, Ocolon, Oda Mari, Ohnoitsjamie, Old Moonraker, Olivier, Omegatron, Onco p53, Opticron, Oreo Priest, Owlbuster, Oxymoron83, Palle.haastrup, Paolo.dL, Papa November, Patrick, Paverider, Pax:Vobiscum, Pdn, Pearle, Pedantik, Pedro, Pekinensis, Penubag, Persian Poet Gal, Pfalstad, Pgadfor, Pgosta, Phasespace, Phil Boswell, Philip Trueman, Photodude, Phynicen, Physis, Pi zza314159, Pinethicket, Pissipo, Pjacobi, Pokipsy76, Pol098, Pollinator, Postglock, Prodego, Qniemiec, Quantpole, Quibik, Quondum, RG2, RJHall, Racko94, RadioFan, Radon210, Rafonseca, Ral315, Razimantv, Rbj, Rdsmith4, Reddi, Regig, RepublicanJacobite, Rich Farmbrough, Richerman, Rico402, Rjstott, Rjwilmsi, Robinh, RockMagnetist, Ross Burgess, Rossami, Rpf, Rrburke, Rundquist, SCZenz, SDC, SJP, Salsb, Salt Yeung, Sam Korn, Santista1982, Sbharris, Sbyrnes321, Scarian, Scohoust, Scooter, Sfu, Shawn81, Sheliak, Signalhead, Sintau.tayua, Sir48, SirEditALot, Sjakkalle, Smack, Smark33021, Smile a While, Sneller2, Snigbrook, Sole Soul, Some jerk on the Internet, SomeUsr, Southen, Speedevil, Srleffler, Ssilvers, Stan Sykora, Stannered, Starwed, Starwiz, Steel, Steve Quinn, SteveBaker, Stevenj, Strait, TStein, Tagray, Talon Artaine, Tarquin, TedPavlic, Telanis, Telpardec, Template namespace initialisation script, Tempodivalse, Tgoyen, Thatguyflint, The Earwig, The Thing That Should Not Be, The wub, The-G-Unit-Boss, Thric3, Tide rolls, Tim Shuba, Tim Starling, Time3000, Tkirkman, Tlabshier, Tonyalfrey, Treisijs, Trovatore, Trusilver, Twsx, UnknownForEver, Useight, Utcursch, Vamaviscool123, Van der Hoorn, Van helsing, Vary, Vcelloho, Veinor, Verdi1, Versus22, Vlus, Wahying, Wavelength, Wavgfkl, Wayward, Whitepaw, WikHead, WikiDao, William Avery, Wizard191, Wogboy52, Wolfkeeper, Woohookitty, Woseph, Xclassmechluv, Ximenes Resende, Xtremepunker, Yevgeny Kats, Yill577, Yoduh2007, Yurei-eggtart, ZodTron, Zoicon5, , , 1064 anonymous edits Magnetization Source: http://en.wikipedia.org/w/index.php?oldid=455385909 Contributors: 16@r, Awickert, BehzadAhmadi, Berserkerus, Brews ohare, Broune US, CambridgeBayWeather, Cdmeyer, D6, David R. Ingham, Emilia.Wiki, Ferengi, Flszen, Fred Hsu, FrozenMan, Gene Nygaard, Giftlite, Grebaldar, Gurch, Harold f, Haymaker, Headbomb, JJ Harrison, Jeff3000, Karol Langner, Kay Dekker, Khalidkhoso, Kjkolb, MER-C, Michael Hardy, Mild Bill Hiccup, Netheril96, Nimur, Pearle, Q0k, RockMagnetist, Rtdrury, Run!, Salsb, San rees, Saperaud, Sbyrnes321, Sitarmooseman, Sugaar, TStein, Tango, TomyDuby, Tone, Ugajin, V8rik, Wdcf, Wolfkeeper, Xenonice, Xxanthippe, Yevgeny Kats, Zhangzhe0101, ^musaz, 46 anonymous edits Magnetic moment Source: http://en.wikipedia.org/w/index.php?oldid=487229671 Contributors: -Majestic-, 122589423KM, Antixt, Anythingyouwant, Arussoxyz, BD2412, Beelaj., Belovedfreak, Bensaccount, Bjheiden, Blastoboy1000, Brews ohare, Can't sleep, clown will eat me, Capricorn42, Charles Matthews, Craig Pemberton, Ctaro, DJIndica, DMacks, Dah31, Daniel.Cardenas, David Shay, Dirac1933, Dreadstar, Dthomsen8, Dusty14, Em3ryguy, Emilio Juanatey, Eric Le Bigot, Ewlyahoocom, F=q(E+v^B), FKLS, Facegarden, Falcon8765, Favonian, Felipe Gonalves Assis, Figure, Freiddie, Fuhghettaboutit, Gab satellite, Gene Nygaard, Geremia, Giftlite, Glosser.ca, Goudzovski, Hadal, IVAN3MAN, Icairns, Icep, Isnow, Jag123, Jim1138, Jozzie0077, Kakila, Kevyn, LeadSongDog, Marx Gomes, MiNombreDeGuerra, Mild Bill Hiccup, Mintleaf, Mwilde, Netheril96, Nick Number, Nmnogueira, Nono64, Oliver Jennrich, Orphan Wiki, Paolo.dL, Pb30, RJFJR, Rex the first, Ricky81682, Robertirwin22, RockMagnetist, Rollred15, Rotiro, Salamurai, Salsb, Sangak, Sankalpdravid, Sbyrnes321, Schmloof, Shadowjams, Sheliak, Skatche, Snigbrook, Steve Quinn, StevenVerstoep, TStein, The Anome, TheQuickBrownFox, Thingg, TomyDuby, Vmenkov, Vokesk, Voyajer, Wavgfkl, WaysToEscape, Wdcf, Whpq, Wikiborg, Wtachi, Wtshymanski, Xenonice, Xxanthippe, Zachlipton, ^musaz, , , 161 anonymous edits Demagnetizing field Source: http://en.wikipedia.org/w/index.php?oldid=476868800 Contributors: 84user, Lmatt, RockMagnetist, Sheepe2004, TStein, Zundark, 2 anonymous edits Magnetic susceptibility Source: http://en.wikipedia.org/w/index.php?oldid=480744846 Contributors: AdjustShift, Apple2, Arc de Ciel, ArsniureDeGallium, Aulis Eskola, BehzadAhmadi, Betacommand, Brews ohare, Capra Ibex Ibex, Chris the speller, Dirac66, DrBob, Duncan.france, Fascinet, Floydmitchell, FrozenMan, Gaius Cornelius, Gcm, Gene Nygaard, Gianluigi, Giftlite, Headbomb, Heron, Icairns, JTB01, Jalanpalmer, Jcwf, Jessetjenkins, Jlacount, Jthorsen3315, Karenjc, LMBRPOVOA, Lightmouse, LilHelpa, Looxix, Lsyn, Lupin, M.O.X, Marie Poise, Michael Hardy, Michi zh, Mlklt, Neparis, Petergans, Politepunk, R'n'B, Radiojon, RahulM18, Rdrosson, RockMagnetist, Roy Hoffman, Salsb, Sbyrnes321, Sciurin, Sebesta, Shanes, Signalhead, Skier Dude, Snigbrook, Tsuji, V1adis1av, Woogee, Wtshymanski, XRiffRaffx, Xenonice, Zureks, , , 69 anonymous edits

Article Sources and Contributors


Permeability (electromagnetism) Source: http://en.wikipedia.org/w/index.php?oldid=485077885 Contributors: 2over0, Aami rony, Aimulti, Andre Engels, Antixt, Archimerged, Army1987, Arnero, Ashishbhatnagar72, Aulis Eskola, BD2412, Barkeep, BehzadAhmadi, Berserkerus, Bmk, Brews ohare, Brockert, Bubba73, Buggi22, Capricorn42, Cdmeyer, Complexica, Conscious, Crazymonkey1123, Cryptic C62, DMacks, Daniel,levine, Deepon, Donarreiskoffer, Draicone, Dxtrous, Ebyabe, Eequor, Electricmic, Evand, Ewilson2011, Ferengi, FrozenMan, Frungi, Fuhghettaboutit, Gene Nygaard, Giftlite, H2g2bob, Hasek is the best, Headbomb, Heron, Hippojazz, Hugo-cs, IanOfNorwich, Icairns, JLD, Jeltz, Jimmy, Jkthomps7, Joechuck, JohnOwens, Jtslm, Kanakukk, Katoa, Keenan Pepper, Keyur86, Kmarinas86, KoenDelaere, L-H, LMB, Lazulilasher, Lovecz, M.O.X, Manco Capac, Marie Poise, Materialscientist, Michael Hardy, Michi zh, MihaiLG, Mmortal03, Nakon, Omegatron, PNG, Pagw, Payakoff, Pearle, PlantTrees, Rdrosson, Rememberway, Rmashhadi, Rtdrury, Ryanrs, Salsb, Sam Hocevar, Sankalpdravid, Sceptre, Shadowlynk, Sibian, Skwa, Snowolf, SocJan, Srleffler, StevenVerstoep, Stevenj, Sundaryourfriend, Sverdrup, TDogg310, TStein, Teapeat, The Land, Thelb4, Tom.Reding, Trevor MacInnis, Verpies, Voidxor, WinstonSmith, Wolfkeeper, Wtshymanski, Xoder, Yoshigev, Z4ngetsu, Zhangzhe0101, Zureks, Zzedar, 144 anonymous edits Force between magnets Source: http://en.wikipedia.org/w/index.php?oldid=488317419 Contributors: Bearcat, Dwayne, Geek1337, GoingBatty, Khazar2, Kmarinas86, Malcolma, NewEnglandYankee, Piast93, Robpparker, RockMagnetist, Siwardio, TStein, Trusilver, 16 anonymous edits Lorentz force Source: http://en.wikipedia.org/w/index.php?oldid=490065323 Contributors: Alexcalamaro, Alfredo, Ambros-aba, Amicus of borg, Ancheta Wis, Andres, BenRG, Boethius65, Brews ohare, Bryan Derksen, CUSENZA Mario, Capricorn42, Chris Howard, Complexica, Conversion script, Cpiral, D-Kuru, D.keenan, DJIndica, DVdm, Deans-nl, Decltype, Dgrant, Dicklyon, DrBob, Drkirkby, Edmundo ba, El C, F=q(E+v^B), FDT, Falcon8765, Frobnitzem, Fuhghettaboutit, FyzixFighter, Gene Nygaard, Geoffrey.landis, George Smyth XI, Giftlite, Headbomb, Heron, HolIgor, Inbamkumar86, InverseHypercube, JNW, JRSpriggs, JabberWok, Jaro.p, Jauhienij, Jcc77, Jdcanfield, Jjalexand, Jkeohane, JohnBlackburne, Jradavenport, K Eliza Coyne, Khazar2, Khunglongcon, Kieff, Kiyabg, Kwamikagami, Laurascudder, Leonard G., Lerdthenerd, Logichulk, Looxix, LtPowers, Lwiniarski, MFNickster, Maschen, Masgatotkaca, Metacomet, Michael C Price, Michael Devore, Mihaip, Mikeblas, Mikiemike, Modeha, Mpatel, Mrdice, Myasuda, Neptune5000, Nick, Nmnogueira, Orderud, Paclopes, Paolo.dL, Petri Krohn, Philip Trueman, Qwasty, Rbj, Reach Out to the Truth, Rich Farmbrough, RobertG, Rror, Rtdrury, Sadi Carnot, Salsb, Sankalpdravid, Sbyrnes321, SebastianHelm, Sfu, Sheliak, Smb1001, Spartaz, StaticGull, SunCreator, Sunnysite, TStein, Tetracube, Tharunsr121, That Guy, From That Show!, The Anome, The mexican boodle, TheBFG, Thurth, Ti89TProgrammer, Tim Shuba, Tim Starling, Treisijs, Tttrung, Uncle Milty, Utcursch, Wavgfkl, Werdna, Wessmaniac, WikHead, Yakeyglee, Yevgeny Kats, , 201 anonymous edits Magnetochemistry Source: http://en.wikipedia.org/w/index.php?oldid=488998771 Contributors: Eumolpo, LilHelpa, Mandarax, Materialscientist, Nadovich, Petergans, Rifleman 82, RockMagnetist, Smartse, 5 anonymous edits Unpaired electron Source: http://en.wikipedia.org/w/index.php?oldid=479080735 Contributors: Benjah-bmm27, Ched Davis, Dpiggy1, Itub, Jcwf, Jia786, Job Inkop, KES47, M-le-mot-dit, Maghnus, Nono64, OlliffeObscurity, Polyamorph, Sai2020, 7 , anonymous edits Atomic orbital Source: http://en.wikipedia.org/w/index.php?oldid=487962152 Contributors: 345gim, A.Z., AManWithNoPlan, AVand, Alansohn, Allstar86, Ambuj.Saxena, Arthur Rubin, Ashandpikachu, Ashmoo, Astrochemist, Audriusa, Avjoska, B7582, BRG, Baccyak4H, Bduke, BeeArkKey, Beetstra, BenB4, BenFrantzDale, Benjah-bmm27, BillyPreset, Black Falcon, Bomac, Bomazi, Bongwarrior, Bowlhover, Br77rino, Bth, CES1596, Calvero JP, Capricorn42, Ceyockey, Chemiker, Christian75, Chzz, Clay Juicer, Cloudswrest, Cmdrjameson, Coffee, Colinphilipjohnstone, Conversion script, Cool3, Cpl Syx, Crystal whacker, Csmallw, DMacks, Ddoherty, Deconstructhis, Delldot, Denisarona, DerHexer, Deror avi, Dirac66, DivineAlpha, Djdaedalus, Djr32, Dlrohrer2003, Donarreiskoffer, Download, Drphilharmonic, EconoPhysicist, Edouard.darchimbaud, Edward, Eequor, Eg-T2g, Ehrenkater, Ekwity, Ema Zee, Enormousdude, Epbr123, EscapingLife, Excirial, Felivik, Fred Bradstadt, Fresheneesz, Fuzzform, Gentgeen, Gerbrant, Giftlite, Gil987, Gilliam, Glacialfox, God Emperor, GoingBatty, Guanaco, H Padleckas, Hairy Dude, HannesJvV, Hans Dunkelberg, Harriv, Headbomb, Hellbus, Helvitica Bold, Heron, HiLo48, Hiyabulldog, HolIgor, Hooperbloob, HorsePunchKid, Hugo-cs, Humanengr, I dream of horses, Ian**, IncognitoErgoSum, Into The Fray, Isaac Dupree, Isilanes, JKW, Jackol, Jan1nad, Janek Kozicki, Jayron32, Jeff02, John Vandenberg, Joo Sousa, Jrockley, Karmalater, Karol Langner, Kmarinas86, Kungfuadam, Kurzon, Kwamikagami, L Kensington, L0ngpar1sh, Laburke, Lightbound, Lightmouse, LilHelpa, LtNOWIS, MER-C, MONGO, Mac Dreamstate, Mako098765, Mani1, Matrix61312, Mattopia, Maurice Carbonaro, Maurreen, Mav, Mayukh iitbombay 2008, Mets501, Michael Hardy, Monfornot, NHRHS2010, NSH001, Nakon, NamfFohyr, Necessary Evil, NeonGas, Neonumbers, Neverquick, Nevit, Notapotato, Ocaasi, OlEnglish, Pagw, Pascal666, Pfalstad, Philip Trueman, Plasmide911, Plotfeat, PrestonH, R'n'B, Raul654, Redrose64, Reelx09, Rich Farmbrough, Rjwilmsi, RobertAustin, RogueNinja, Romanm, Ronhjones, Rotational, SDC, SUL, Sadi Carnot, Samwb123, Sbharris, Scerri, Schoen, SeventyThree, Shirik, Simstud16, Slakr, Slightsmile, Slon02, Smack, Smokyhallow, Snobscure, Spiral5800, Stevey7788, Stismail, StradivariusTV, THEN WHO WAS PHONE?, Tardis, Tarquin, Taxman, TelecomNut, Tesi1700, The Master of Mayhem, The Thing That Should Not Be, Thrindel, Tide rolls, Tim Starling, Tomas e, Tonyrex, Topbanana, Trevorhailey1, Trigley, Ttony21, User A1, V8rik, VanishedUser314159, Velella, Vinay.bhat, Virginia fried chicken, Voyajer, Vsmith, Wavelength, Wayne Slam, Wd930, Whoop whoop pull up, Wikiphysicsgr, William M. Connolley, WinterSpw, Xetrov znt, Yoshigev, 312 anonymous edits d electron count Source: http://en.wikipedia.org/w/index.php?oldid=446656460 Contributors: Chris the speller, Dirac66, Gaius Cornelius, Giladbr, Hellbus, OMCV, Quadell, ZooFari, 11 anonymous edits Hund's rule of maximum multiplicity Source: http://en.wikipedia.org/w/index.php?oldid=477035128 Contributors: Acegikmo1, Allen3, Bluetooth954, Bunnyhop11, Chrislk02, Csigabi, DMacks, Diannaa, Dirac66, Dr.K., Edsanville, Flying Jazz, Gentgeen, I, Englishman, Icairns, Imnotoneofyou, Itub, Jrockley, Karada, Karol Langner, Keenan Pepper, La goutte de pluie, Llwrce, Mekong Bluesman, Mormegil, NuclearWarfare, Nuklear, Orborde, Peko, Philip Trueman, Pietaster, Proski, Ram einstein, Rcsprinter123, Sadi Carnot, Scientific29, Stemonitis, Stevey7788, Theresa knott, Tide rolls, Vespristiano, XinaNicole, Ziel, , 52 anonymous edits Aufbau principle Source: http://en.wikipedia.org/w/index.php?oldid=479483425 Contributors: A. di M., Aaleo20, Amit kumar roy, Andres, Andrew Gray, Arthur Rubin, Arvindsrm2894, BWDuncan, Bobo192, Bschaeffer, Btharper1221, CardinalDan, Chemmix, DMacks, Dirac66, Divye, DocWatson42, DragonflySixtyseven, Drova, ErkangZhu, Escherichia coli, Ettrig, Firsfron, Gershwinrb, Graham87, Guillom, Hans Dunkelberg, Ianprime0509, Ibt, Itub, Joker99352, Jon the Geek, Jrockley, Julia Neumann, Killing Vector, Lanthanum-138, Laurinavicius, Meatmanek, Mgiganteus1, Mitosh mora, Moonraker12, Neelix, OwenBlacker, PJTraill, Philip Trueman, Physchim62, Pieter Kuiper, Pietz, Pluma, Postglock, Pozitron969, Prashanthns, R8R Gtrs, Rettetast, Rice.brendan, Rifleman 82, Robofish, Rundquist, Shureg, SidP, SimDarthMaul, SimonP, Smack, Some jerk on the Internet, Spaghettipoop, Spoon!, Stefano85, T.vanschaik, Tesi1700, Tomj, User A1, Versus22, Wang lvan, Zaheen, 28 , anonymous edits Coordination complex Source: http://en.wikipedia.org/w/index.php?oldid=489105273 Contributors: 4lex, Alexsanjoseph, Algebraist, Alvestrand, Arzoo jain, Aushulz, BTDenyer, Beetstra, Benjah-bmm27, Bhellis, Bigbluefish, Bomac, Borgx, Brane.Blokar, Brendio, Cacycle, Cadmium, Centrx, Chiu frederick, Chrumps, Cmrufo, Conversion script, Crystal whacker, DMacks, DVirus101, David H Braun (1964), Dcoetzee, Delirium, Dirac66, Download, Drphilharmonic, Dslate2123, Dwmyers, Eg-T2g, ElZarco, Essin, Eumolpo, Feneeth of Borg, Freestyle-69, Fuyutsuki, Galwhaa, Gentgeen, Geoffreybernardo, Glenn, Hellbus, Hmrox, Hugo-cs, Innotata, Itub, Jasper Deng, Jeffq, Jimfbleak, Johner, Jubin256, Kehenr01, Loupeter, Malbi, Man of Meat, Matthias M., Mboverload, Michael Hardy, Miniyazz, NSR, Nono64, Nunh-huh, OMCV, Ojigiri, Olin, Oxymoron83, Petergans, PierreAbbat, Pjacobi, Pjetter, Polyamorph, Postglock, Psiphiorg, RainbowOfLight, Richard.decal, Rifleman 82, Rjwilmsi, Sac5sc, ScAvenger, SeventyThree, Shalom Yechiel, Smokefoot, Sodium, Steeev, Stepa, Steve Quinn, SteveLower, TangParadise, TheTito, Tom harrison, Uncle Dick, V8rik, Violentbob, Viswaprabha, Walkerma, Whmice, Wickey-nl, Wierdw123, Xiaopo, Yahya Abdal-Aziz, Yath, 177 anonymous edits Diamagnetism Source: http://en.wikipedia.org/w/index.php?oldid=489936344 Contributors: 213.253.39.xxx, 24.1.200.xxx, AJim, ALACE, Aaagmnr, Acroterion, Adashiel, Arkadipta banerjee, Bakuryuu, Beland, Belg4mit, Bluefalcon07, Bodnotbod, Brews ohare, Bryan Derksen, Busukxuan, Campuzano85, Candleknight, Casey boy, Cesiumfrog, CharlesC, Cheeseifyer, Constructive editor, Conversion script, Cp111, Cquan, DARTH SIDIOUS 2, Darekun, DarkHorse, Deepnightblue, Deglr6328, Dfinkel, Dimwitt Flathead, Dirac1933, Don4of4, DragonflySixtyseven, Dwmyers, EbedYahweh, Eigenpirate, Favonian, Foobar, Gaius Cornelius, Gene Nygaard, Georgelazenby, Giftlite, Glacialfox, Graham87, Gudeldar, Guswandhi, Hans Dunkelberg, Headbomb, Hede2000, Henrygb, Heron, Hesperian, Horkana, Icairns, Iliev, Inter rest, Jaapkroe, Jackelfive, Jafet, Jcline1, Jcwf, Jinxed, Jkeohane, Joanjoc, JustAddPeter, K Eliza Coyne, Kaifeng, Karol Langner, KasugaHuang, Katalaveno, Kmarinas86, L'Aquatique, Leobh, Lfh, LogaRhythm, Looxix, Lumrs, Macderv15h, Mbweissman, Mech Aaron, Midgrid, Mike Rosoft, MisterSheik, Mmm, Moemin05, Netscott, OlEnglish, Oli Filth, Omegatron, Pearle, Peterburton, Petergans, Pharaoh of the Wizards, Phoenix79, Phys, Piil, Planetscared, Poisonmilk, Prikryl, Rage, Rememberway, Rifleman 82, Roadrunner, Robin Whittle, RockMagnetist, Salsb, Sappe, Scott Dial, Serverxeon, Sibian, Sikkema, SilentOpen, Silly rabbit, Slakr, Smalljim, Smokefoot, Snigbrook, Splarka, Stokerm, Suffusion of Yellow, Tarotcards, TedPavlic, Tim Starling, Tmadge, Tomothy, Troyrock, Vanderdecken, Vanished user, Vrenator, Vsmith, WLU, Waleswatcher, Whitepaw, Wolfkeeper, Xanzzibar, Xompanthy, Yakiniku, Zamirm, Zereshk, Zinger0, , , 182 anonymous edits Paramagnetism Source: http://en.wikipedia.org/w/index.php?oldid=482806554 Contributors: Aaagmnr, Admiral Norton, Ahoerstemeier, AlexGWU, Anchananatarajan, Andre Engels, Art and Muscle, Bandy, Bduke, Beetstra, Benbest, Berland, Brews ohare, Busukxuan, CRON, Cmcnicoll, Complexica, Conversion script, Cyanoir, DanielRigal, Dcoetzee, DeadEyeArrow, Dina, Eborreson, Electricmic, Electron9, Eric Kvaalen, F-402, Freddy78, Gadolinist, Gamera2, Gene Nygaard, Giftlite, Hammersbach, Heron, HopeChrist, Humanist, Icairns, Inquisitus, IronGargoyle, IvanLanin, Jcwf, Jdedmond, Joanjoc, John, JorisvS, KathrynLybarger, Kbrose, Koweja, Lfh, Light current, Looxix, Ls1955, Luckas Blade, Lugh23, Mac Davis, Markjoseph125, Materialscientist, Mdsam2, Medeis, Mercurywoodrose, Mgiganteus1, Michael Hardy, Murukesh mohanan, Mwhiz, NCurse, Nick Pisarro, Jr., No1lakersfan, Omegatron, Oxymoron83, Pde, Peter.C, Petergans, Phys, Piano non troppo, Piil, Polyamorph, Potatoswatter, Ppxatc, Profero, Quantum7, Quibik, Rakista, Rangergordon, Rebroad, Rifleman 82, Robinsoncrusoe, RockMagnetist, RocketDavid, Salsb, Samaritan13, Sbharris, Schneelocke, Silenced, Silverplasma, SimonArlott, Slakr, Smokefoot, Someones life, Sonicology, Stevenj, Stokerm, Tassedethe, Taw, Thevenin77, Thumperward, Tim Starling, Troyrock, Turbos10, WhiteDragon, Zero sharp, 871 ,anonymous edits Electron magnetic dipole moment Source: http://en.wikipedia.org/w/index.php?oldid=489931586 Contributors: Agesworth, Altenmann, Anythingyouwant, Appleseed, BD2412, Bookalign, Brichcja, Chibibrain, Dino, Djjew, Dominic, F=q(E+v^B), Gavoth, Giftlite, GoingBatty, Grj23, HEL, Headbomb, Kevmitch, Khazar2, Kusma, LAk loho, Linas, Little.pig.microphone, Maliz, Mohawkjohn, Mowerm, Niout, Pagw, Pearle, Phelimb, QFT, RJHall, Sbyrnes321, Voyajer, Wavelength, XJamRastafire, 19 anonymous edits

254

Article Sources and Contributors


Pascal's constants Source: http://en.wikipedia.org/w/index.php?oldid=416046517 Contributors: Christian75, Petergans, Smokefoot, 1 anonymous edits CurieWeiss law Source: http://en.wikipedia.org/w/index.php?oldid=457242561 Contributors: Aeronautics, BrokenSegue, Gene Nygaard, HHahn, Jheald, Juanjovsky, Kropotkine 113, Latifahphysics, Legoktm, Michael Hardy, Mpfiz, RockMagnetist, Salsb, Tegla, TomViza, Tone, V1adis1av, Yevgeny Kats, YuryKirienko, 5 anonymous edits Curie's law Source: http://en.wikipedia.org/w/index.php?oldid=475886522 Contributors: A.shteiman, Antixt, Beaber, CBM, CapitalR, Choster, Ddcampayo, E0steven, Gene Nygaard, Giftlite, Ideal gas equation, Jag123, Macholl, Mbell, Mercenario97, Mnmngb, OlEnglish, PV=nRT, Piil, PloniAlmoni, RDBury, RJFJR, Rror, Salsb, Sbyrnes321, SemperBlotto, Thatdog, The wub, Tizeff, Unc.hbar, UncleDouggie, 38 anonymous edits Curie constant Source: http://en.wikipedia.org/w/index.php?oldid=486059212 Contributors: Alansohn, Brammers, Choihei, Cjsmed, Elm-39, Giftlite, Goldenrowley, JHobbs103, Michbich, Parusaro, Piil, Proofreader77, Pyrosim, SteinbDJ, Talso chui, Tempodivalse, Tone, 34 anonymous edits Boltzmann distribution Source: http://en.wikipedia.org/w/index.php?oldid=489117810 Contributors: Anoko moonlight, Btyner, Conversion script, Csigabi, Danski14, Dicklyon, Enormousdude, Eranb, Geepster, Giftlite, Graham87, Grimlock, Halberdo, Headbomb, Isopropyl, Jheald, Kmarinas86, Linas, Lostart, Luke R001, Mct mht, Michael Hardy, Mnmngb, Od Mishehu, Oolongy, PAR, Pavdpr, Qwfp, Salix alba, Sho Uemura, Shreevatsa, Stepa, Stokerm, Thedeejay, Wisebridge, Zainalarifin2028, 20 anonymous edits Brillouin and Langevin functions Source: http://en.wikipedia.org/w/index.php?oldid=483993707 Contributors: Antixt, Azhyd, Bbanerje, Bcartolo, Bender235, Bomazi, CBM, Ddcampayo, Edgar.bonet, Emersoni, Fabrictramp, Flowerpotman, HappyCamper, Harmoxe, Headbomb, Sbyrnes321, Snaful, Tango, 9 anonymous edits Gyromagnetic ratio Source: http://en.wikipedia.org/w/index.php?oldid=487310195 Contributors: Alinaschimpf, Chris the speller, Chronofied, Constructive editor, Cuaxdon, Devon Fyson, Dirac1933, Edoarado, Elroch, Fantumphool, Fernando.Martin.Maroto, F, GTBacchus, Gadzirai, Gene Nygaard, Gillen, Gonfer, Goudzovski, Grammaticus Repairo, Gyro Copter, Headbomb, Jag123, John Vandenberg, Linas, Markaci, MicahJonson, Mnmngb, Nick Mks, Nkuzmik, Oerms, Pegship, Plastikspork, Rjwilmsi, Robin Kerrison, Rotiro, Sbyrnes321, Seaphoto, SeventyThree, Sheepe2004, Spiffyzha, Suiy02, TomyDuby, Tone, Voyajer, Wiki me, Wolfmankurd, Xenonice, Xzapro4, Zas2000, 60 anonymous edits Electron paramagnetic resonance Source: http://en.wikipedia.org/w/index.php?oldid=487739130 Contributors: Ahheiss, Astrochemist, Bci2, Brewhaha@edmc.net, Brichcja, Bruker, Chem-awb, Cybercobra, DarkArcher25, Der Marc, Drphilharmonic, EDC105, Ehotze, Erichpi1, Falconflyer94, Fisik 83, GavinMorley, Gene Nygaard, Greyhood, Jauhienij, Jcwf, Jotomicron, K Eliza Coyne, Kbk, Kivirus, Kkmurray, LilHelpa, LouisBB, Materialscientist, Mikhail Ryazanov, Nafradi, Oxymoron83, Pavel Dusek, Pradameinhoff, Rjwilmsi, Sbyrnes321, Shiroi Hane, Simetrical, Skier Dude, Spectro Sci, Stan Sykora, StaticGull, Stephen.c.johnson, Tarotcards, Tetracube, Thaejas, Transmobilator, Untifler, V8rik, Van helsing, Wtmitchell, Xenonice, Yuckfoo, 17 , anonymous edits Larmor precession Source: http://en.wikipedia.org/w/index.php?oldid=485310420 Contributors: Bauleaf, Berland, Charvest, Choij, Chutznik, Csigabi, Danski14, Dchristle, Dfeldmann, Dirac66, G-W, Gerasime, Gillen, Henry Delforn (old), Karol Langner, Linas, Michael Hardy, Mnmngb, Nephron, Nyctea, Pegship, RekishiEJ, RockMagnetist, SciYann, Stone, Tdonoughue, ThePI, Tormance, WISo, Yamavu, 29 anonymous edits Pulsed EPR Source: http://en.wikipedia.org/w/index.php?oldid=452107751 Contributors: Erichpi1, GavinMorley, HBook, Materialscientist, Mild Bill Hiccup, RadioFan, 2 anonymous edits Bohr magneton Source: http://en.wikipedia.org/w/index.php?oldid=483749261 Contributors: 478jjjz, Afil, Altenmann, Antikon, Anythingyouwant, Bci2, Complexica, Cristi Stoica, Danski14, David C Bailey, Evgeny, Gene Nygaard, Headbomb, Ideal gas equation, Isnow, Jimp, Jonathan A Jones, Karol Langner, Kevmitch, KnightRider, Korech, Linas, Lone F, Looxix, Magnus Manske, Malbi, Martin Ulfvik, Muhends, Owlbuster, Pagw, Pieter Kuiper, Rickproser, Salsb, Sbyrnes321, Semolo75, Spamhunt, Tkynerd, Van helsing, Vlado4, Wenboown, Youandme, 33 anonymous edits Spin (physics) Source: http://en.wikipedia.org/w/index.php?oldid=490048213 Contributors: 4lex, 8af4bf06611c, A. di M., A876, Accordionman, Aepryus, Ahmes, AlistairMcMillan, Andre Engels, Andrei Stroe, Andres, Andycjp, Anonymous Dissident, Ant59, Anthony, Army1987, Ashmoo, Astavats, Ato, Atomicann, AugPi, Average Earthman, AxelBoldt, Bambaiah, Barticus88, Beatnik8983, Beefyt, Beland, Bevo, Bidabadi, Blainster, Bmk, Bo Jacoby, Bolzano, Brichcja, Brilliand, Buster79, C.Bluck, CRGreathouse, CYD, CardinalDan, Casimir9999, Charles Matthews, Cmichael, Complexica, Costela, Count Iblis, Craig Pemberton, Craklyn, Crowsnest, Cyp, DHN, DMacks, DVD R W, DVdm, Daniel.Cardenas, David Trochos, David spector, Davidhorman, Diberri, Djdaedalus, Dominus, Dreftymac, Drrngrvy, E23, Eg-T2g, Elroch, Em3ryguy, Eric Le Bigot, Ernsts, Error9312, Ewlyahoocom, F=q(E+v^B), FilipeS, Francine Rogers, Frau Holle, Freddy78, Fresheneesz, Gandalf61, Garuh knight, Giftlite, Glenn, Gurch, Gzuckier, Hadal, Hairy Dude, Harddk, Headbomb, Hqb, Hwasungmars, Igodard, Igorivanov, Inquisitus, Itinerant1, J.delanoy, JRSpriggs, JabberWok, JackSchmidt, Jacksccsi, Jaganath, Jaksmata, Jasondet, Jgrahamc, Jheise, Jkforde, JohnBlackburne, Jonas Ferry, Jondn, JonnybrotherJr, Joostvandeputte, Joshronsen, KSevcik, Kallikanzarid, Karol Langner, KasugaHuang, Kbk, Kbrose, Keenan Pepper, Kganjam, Kri, Larry Sanger, Lendtuffz, Leo C Stein, Lethe, Lightmouse, Linas, Loodog, LorenzoB, Lupin, MER-C, MathKnight, Mathieu Perrin, Maurice Carbonaro, Mct mht, Med, Melongrower, MichaelBillington, Mike Rosoft, Mild Bill Hiccup, Miracle Pen, Monurkar, Mpjohans, MuDavid, Mxn, Myasuda, NawlinWiki, Neh0000, Neonumbers, Nepahwin, Netrapt, Newone, Nikai, NoldorinElf, Numbo3, Obradovic Goran, Oreo Priest, Orionus, Paine Ellsworth, Paolo.dL, Papadopc, Paul August, Peachypoh, Petergans, PhiMAP, PhilKnight, Phys, Physicistjedi, Pokyrek, Pradameinhoff, Pt, QFT, R'n'B, Radagast83, RadicalOne, Rasinj, Reina riemann, Richard001, Rjwilmsi, Robertwilliams2011, RokasT, Rossami, Ryanluck, Sackm, Sadi Carnot, Sangak, Sbian, Sbyrnes321, Sergio.ballestrero, Sgravn, Shanesan, Simen, Sionus, Smjg, SpeedyGonsales, Spiralhighway, Sriharsha.karnati, Srleffler, Stevenj, Stevertigo, Susvolans, Swpb, TechnoFaye, Terry Bollinger, Terryeo, Tesseran, That Guy, From That Show!, The Anome, The way, the truth, and the light, TheMidnighters, Tide rolls, Tim Starling, Timwi, Tlabshier, Tobby72, Tom1936, Torquil, Tsemii, UberCryxic, Uberdude85, V1adis1av, Vgy7ujm, Voyajer, WandringMinstrel, Warbler271, Werdna, Wierdw123, Wj32, Wood Thrush, Wtrmute, XJamRastafire, Xavic69, Xerxes314, Yudem, Zack, Zhenyok 1, Zoz, 281 anonymous edits Land g-factor Source: http://en.wikipedia.org/w/index.php?oldid=435917001 Contributors: ArkadiuszBokowy, Charland, Charles Matthews, Dchristle, Discospinster, DragonflySixtyseven, Gillen, Linas, Mako098765, Mathieu Perrin, Sbyrnes321, Schmloof, Tobias Bergemann, Voyajer, Wiki me, Zebas, ^musaz, 15 anonymous edits g-factor (physics) Source: http://en.wikipedia.org/w/index.php?oldid=471947384 Contributors: A. di M., AManWithNoPlan, Army1987, BjoernZ, Brews ohare, Brichcja, Cgd8d, Crowsnest, Doug, Ettrig, Fwb22, GTBacchus, Gene Nygaard, Giggy, Gwern, Headbomb, John Vandenberg, Lilac Soul, MER-C, Mako098765, Maliz, Nvj, Potekhin, RJHall, Rjwilmsi, Sbharris, Sbyrnes321, User A1, Victor Chmara, Wiki me, Xenonice, Ybband, 33 anonymous edits Zeeman effect Source: http://en.wikipedia.org/w/index.php?oldid=488817383 Contributors: .:Ajvol:., A. di M., Adiel lo, Andrew c, Arjayay, AstroNomer, Babbage, Bogdangiusca, Brian0918, Caid Raspa, Casull, ChrisGualtieri, Chuunen Baka, Columbiafan, Cyrius, Danski14, Dboonz, Deflective, Denelson83, Dhanak, Difameo, Djegan, Dukeofalba, Ehinson56, Eric Le Bigot, Escape Orbit, Evgeny, Evlekis, Fradeve11, FreakingOut, FreplySpang, Geboy, Georgelulu, Giftlite, Gregors, HEL, Heron, HiraV, Iantresman, JabberWok, Jaganath, Jerry-VA, Jora0, Karol Langner, Keenan Pepper, Kwamikagami, Lapsus Linguae, Latifahphysics, Laurascudder, Linas, Lissajous, Looxix, Magnus Manske, MicioGeremia, Mike Peel, Milanmart, Mnmngb, MortimerCat, MyFaJoArCo, Myasuda, Nick Mks, Nigholith, Nsh, Oasisbob, Orioane, Ozguroot, Pagw, Qwertyus, Rjwilmsi, RockMagnetist, Salamurai, Salsb, Sefzik, Shoeofdeath, Smite-Meister, Some jerk on the Internet, Steven Zhang, That Guy, From That Show!, Tim Starling, Tomkoltai, Tsemii, Txomin, Voyajer, WMdeMuynck, Xnchu, Youandme, Ysangkok, Zarniwoot, 115 anonymous edits Spin states (d electrons) Source: http://en.wikipedia.org/w/index.php?oldid=478768849 Contributors: John of Reading, Kmarinas86, NameIsRon, OMCV, Petergans, Prari, Quadell, V8rik, 15 anonymous edits Spinorbit interaction Source: http://en.wikipedia.org/w/index.php?oldid=489304452 Contributors: A. di M., Alex-engraver, AlexandrF, BWDuncan, Charvest, Chibibrain, DJIndica, Dan Gluck, Daniel.Cardenas, Droll, Elpaw, Ethereal-Blade, Franco3450, Gaius Cornelius, Gerasime, Gillis, Grimlock, HEL, Harddk, Hpubliclibrary, Kevmitch, Linas, Loodog, LorenzoB, Mets501, Mh26, Mushin, NULL, Paradoctor, Radagast83, RockMagnetist, Sbyrnes321, Schmloof, Snezak, Sunilsrivastava, T-borg, That Guy, From That Show!, Thurth, Yquiensabe, Zylorian, Zzedar, 59 anonymous edits Azimuthal quantum number Source: http://en.wikipedia.org/w/index.php?oldid=486701537 Contributors: Angusmca, Antixt, Barticus88, Betacommand, Bloodoflamb, Brittl33, Chuunen Baka, Craig Pemberton, Dan Gluck, DomenicDenicola, Donarreiskoffer, Edobson, Edward Z. Yang, Elronxenu, Freiddie, Fresheneesz, Giftlite, Headbomb, Heptazane, Hooperbloob, Ian**, Incnis Mrsi, Joyonicity, Kallikanzarid, Karada, Kborland, Keflavich, Larryisgood, Latch.r, Linas, Lus Felipe Braga, Magicxcian, Maurice Carbonaro, Meddlingwithfire, Mni9791, Mxn, P.wormer, Radagast83, Rich257, Rjwilmsi, Robert Weemeyer, Rvodden, Sbyrnes321, Shal Dengeki, Sohum, Spoon!, Stephan Leeds, Superplaya, TZGreat, TeleComNasSprVen, Tom Lougheed, Tomisti, Venny85, Voyajer, Vsmith, Whitepaw, William Avery, Zscout370, 61 anonymous edits Principal quantum number Source: http://en.wikipedia.org/w/index.php?oldid=489793331 Contributors: Abhishek anand2208, Aeluwas, Arkenflame, Astrochemist, Benjah-bmm27, Blumschein, Cayte, Conscious, Der Falke, Donarreiskoffer, Edward Z. Yang, Escape Orbit, Giftlite, Heptazane, I c dead phat homies, Ian**, Itub, Jack B108, KyuubiSeal, Larryisgood, Lifeisaserenade, Linas, Lus Felipe Braga, Michael Hardy, Mni9791, Mxn, Ninetyone, PigAndPepper, Roberdin, Sammymiceli, Shanes, SiriusB, Spiritia, The Anome, Tom Lougheed, Tomisti, UnPiccoloPrincipe, Veryangrypenguin, Voyajer, WDavis1911, Wayne Slam, Whosyourjudas, William Avery, Zapateria, 47 anonymous edits Spin quantum number Source: http://en.wikipedia.org/w/index.php?oldid=487512804 Contributors: Afa86, Bobo192, Carlog3, Chris the speller, Daniel.Cardenas, Dark Formal, Djdaedalus, Dominus, Drphilharmonic, Edward Z. Yang, Fastfission, Favonian, Gary King, GavinMorley, Giftlite, Heptazane, Hermitage17, Incnis Mrsi, Itinerant1, Jochietoch, Latch.r, LaudanumCoda, Linas, Marceluda, Mbell, Mxn, Myasuda, Obscurans, Radagast83, Ruislick0, Sango123, Tlabshier, Tom Lougheed, Tomisti, Voyajer, Whosyourjudas, Writtenright, Xavic69, 53 anonymous edits

255

Article Sources and Contributors


Total angular momentum quantum number Source: http://en.wikipedia.org/w/index.php?oldid=486069800 Contributors: Antixt, Brews ohare, Dan Gluck, Duckyphysics, Gillen, Guguma5, Incnis Mrsi, JabberWok, JohnBlackburne, Linas, Mbell, Michael Hardy, Newty23125, That Guy, From That Show!, Tim Starling, Xavic69, 4 anonymous edits Angular momentum operator Source: http://en.wikipedia.org/w/index.php?oldid=486832002 Contributors: Algebraist, Circeus, Ettrig, F=q(E+v^B), Georgelulu, INVERTED, JRSpriggs, JohnBlackburne, Jvgama, Kallikanzarid, KathrynLybarger, Kborland, Linas, Michael Hardy, MuDavid, Netheril96, Nielswalet, Niout, Oleg Alexandrov, Rjwilmsi, Sbyrnes321, Substar, Thurth, Yngvadottir, 18 anonymous edits Angular momentum Source: http://en.wikipedia.org/w/index.php?oldid=489293233 Contributors: .:Ajvol:., 05daygeo, 165.123.179.xxx, 6birc, AManWithNoPlan, Abhishek727, Adam Rock, Alan Smithee 87, Alansohn, AlexBrett, Andonee, Andrei Stroe, Andries, Anonymous Dissident, Arbitrarily0, ArglebargleIV, Army1987, Asplace, AstroNomer, AugPi, AxelBoldt, Azuring, Barticus88, BenFrantzDale, Bensaccount, Bo Jacoby, Bowlhover, Brews ohare, Buster79, CambridgeBayWeather, Capecodeph, CarlosPatio, Charles Matthews, Chetvorno, Closedmouth, Coffee, Complexica, Conversion script, Crazynas, DCEdwards1966, DMacks, Dan Gluck, Daniel.Cardenas, Dark Formal, Darthmgh, Dave6, Davidryan168, Dedicated to learn, DerHexer, Dicklyon, Dkasak, Dolphin51, Donatus, DrBob, EconoPhysicist, Eigenlambda, El C, Enochlau, Enormousdude, F3meyer, Fashionslide, Fibonacci, FilipeS, Fourier Transform, Fresheneesz, Fuhghettaboutit, Gaius Cornelius, Gene Nygaard, Giftlite, Glenn, GliderMaven, Gonzalo Diethelm, H Padleckas, H2g2bob, Happy-melon, Headbomb, Heron, Hooperbloob, Ian Glenn, Icairns, Iiar, Incnis Mrsi, J04n, JCO312, JaGa, JabberWok, JakeVortex, JeffBobFrank, Jimp, Jj1236, Jmcc150, Joanjoc, JohnOwens, Jonkerz, Jorfer, Joshua Issac, Karol Langner, Kenny56, Kongr43gpen, Kristen Eriksen, Ksant3, Laoma, Laterensis, LaurelESH, Lfstevens, Linas, Lmarrucci, Localzuk, Loodog, Looxix, LorenzoB, Lowellian, Lseixas, Maelnuneb, Magioladitis, MarcusMaximus, Maschen, MathKnight, Matusz, Maurice Carbonaro, Melchoir, Melongrower, Mets501, Michael Hardy, Mike2vil, Milez, Monurkar, Mr Stephen, MuDavid, NerdyScienceDude, Nikai, Norvy, Oda Mari, OldManNIck, Out of Phase User, Oxymoron83, P.wormer, PAR, Passw0rd, Patrick, Paul August, Pdefer, Pharaoh of the Wizards, Phil Holmes, PhySusie, Phys, Pkbharti, Psymun747, Quibik, Qutezuce, R3m0t, Radon210, Rednblu, Rich Farmbrough, Richard001, Rjwilmsi, Rmhermen, Ro234, RobertHannah89, RobertRHannah89, Rogper, Rohitphy, Ronhjones, Rossami, Rpiphysicsnerd1111, Rrburke, Sanpaz, Savant13, Sbyrnes321, Scintz, Selkem, Shalom Yechiel, Shivsagardharam, Solvang, Special-T, Sreyan, Srleffler, Sselig, Starwiz, Steve Quinn, StewartMH, StradivariusTV, Taichi, Tarquin, Teapeat, TeeEmCee, Tesi1700, The Anome, The Thing That Should Not Be, TheAMmollusc, TimothyRias, Timsdad, User A1, Vonregensburg, W0lfie, Wavelength, WaysToEscape, Wolfkeeper, Xanzzibar, Xenonice, Xxanthippe, Yawe, Zaidpjd, Zdtan, Zero sharp, Zueignung, Zzigee, 322 anonymous edits Magnetic quantum number Source: http://en.wikipedia.org/w/index.php?oldid=480659478 Contributors: Bloodoflamb, Bongwarrior, Dark Formal, Donarreiskoffer, Edward Z. Yang, ElTchanggo, Giftlite, Gurch, Headbomb, Heptazane, Icep, JamesAM, Larryisgood, Linas, Mxn, Nebarnix, NoobOfWar, Paolo.dL, Pearle, PhilKnight, Purgatory Fubar, Ram einstein, Redrose64, Rettetast, Spoon!, Stone, Thryduulf, Tom Lougheed, Tomisti, V1adis1av, Voyajer, 29 anonymous edits Pauli exclusion principle Source: http://en.wikipedia.org/w/index.php?oldid=487357361 Contributors: 128.12.93.xxx, AgadaUrbanit, Ahoerstemeier, Alan Peakall, Alexander1102, Ambuj.Saxena, Andre Engels, Andrei Stroe, Andres, Animum, Anonymous Dissident, Army1987, Aryabhatta, BTDenyer, Beeline23, BenFrantzDale, Brews ohare, Bryangv, CYD, Canice, Chamal N, Charles Matthews, CharlotteWebb, Chenyu, Cherlin, Chief buffalo chip, Cognitivelydissonant, Complexica, Conversion script, CosineKitty, Craig Pemberton, Crucis, Custos0, D Hill, DJIndica, DVdm, David G Brault, Dirac66, Dlabtot, Dominus, Dvorsky, EddEdmondson, ElCharismo, ElusiveByte, Emperorbma, Enochlau, Enormousdude, Enzo Aquarius, Ethereal-Blade, Eyu100, FT2, Fibonacci, FourBlades, Fresheneesz, Georgelulu, Giftlite, Gioto, Glenn, Glennwells, Gombang, Graham Chapman, Graham87, Graue, Greg L, Gromonger-17, Gunnar Larsson, Gurevichar, Harp, Headbomb, Heisenbergthechemist, Henry Delforn (old), Hickorybark, HorsePunchKid, IEROslippersBRYAR, IW.HG, Iain99, Icek, Isnow, Iwilcox, JRSpriggs, JabberWok, Jakeng, Jehochman, Jheise, John Darrow, John Vandenberg, JohnMarkStrain, Karol Langner, Kaszeta, Katalaveno, Keenman76, Korepin, Kriak, Krishnaprasaths, LeaveSleaves, Liberatus, Likebox, Materialscientist, Mattbr, Maximum bobby, Meco, Mejor Los Indios, Melamed katz, Michael C Price, Michael Hardy, Michalchik, Modeha, Montrealais, Mrfunkyostrich, Mrhsj, Nathanielvirgo, Nickkid5, Obradovic Goran, Okedem, Orionus, Otis182, Owain, P.wormer, Phr en, PierreAbbat, Pishogue, Pol098, Pseudospin, Quaint and curious, Quondum, RG2, RJC, RJFJR, RJHall, RQG, Ravindraia, Relke, RexNL, Riick, Ripper234, Roadrunner, Sadi Carnot, Shai-kun, Sheliak, Shellreef, Smalljim, Sohum, Srleffler, Stevenj, Stevertigo, Strami, T.O. Rainy Day, Tanvirzaman, That Guy, From That Show!, Thingg, Thorseth, Tim Starling, Tjlafave, Tobias Bergemann, TristramBrelstaff, Trojancowboy, Trugster, Tsemii, UnitedStatesian, Voltagedrop, Vyznev Xnebara, Wasell, Wavelength, Wfructose, WhiteHatLurker, Wiki alf, Wikiborg, XJamRastafire, Xcentaur, Zipz0p, Zoedill, 192 anonymous edits Ferrofluid Source: http://en.wikipedia.org/w/index.php?oldid=489529530 Contributors: Abiermans, AlistairMcMillan, Anaxial, Andres, ArielGold, Arjayay, ArmadniGeneral, AzaToth, BananaFiend, Bando26, Binksternet, Bobblewik, Bobo192, Bookofjude, Brian the Editor, Cardamon, Ceyockey, Chachilongbow, DVdm, David Newell, Dawidl, Debresser, Denisarona, Dgw, Dhollm, Dhwanit.zaveri, Dirac1933, Donohuen, DragonflySixtyseven, Droll, Dzubint, Edgewise, Edusilva, Electricmic, Esmhead, Eyrian, Fama Clamosa, Ferrofluidtoy, Frap, FumarMata, GeeJo, Geniac, Giancarlo Rossi, Giraffedata, Gliktch, Glsps, Gmaxwell, GregorB, GroupeBorra, Harlekeyn, Harold f, Headbomb, Heron, Hetar, Hon-3s-T, Idran, ImZenith, Inferno, Lord of Penguins, Infinityfsho, InversionControl, IstvanWolf, J.delanoy, Jarfil, Jbusenitz, Jcwf, Jeepday, JorisvS, Josh.karli, Julia W, Jrme, Katalaveno, Kcordina, Krishnavedala, KurtRaschke, Lawl Rock, Light current, Lightmouse, Lofty, Mac Davis, Marie Poise, Materialscientist, Mboverload, Melchoir, Michaelas10, Mikiemike, Mnolf, Moshe Constantine Hassan Al-Silverburg, Mr Beige, MrBell, Mxn, Naff89, NapoliRoma, Narayanarora, Nauticashades, Nigosh, ONUnicorn, Ocaasi, Omegatron, Opoterser, Ouro, ParisianBlade, Pauwel, Philip john123, Phoenix-forgotten, Pinestone, Plasticup, Poeloq, Ppxatc, Prari, Puchiko, RainbowOfLight, Rees11, Rich Farmbrough, Rjwilmsi, Roadmr, Roulianne, Ryulong, S2000magician, Salsb, Savant13, Sfoskett, Shaddack, Shanedidona, Shoeofdeath, Shrampes, SithiR, Sjb90, Skierpage, Sonoma-rich, SteveBaker, Synethos, Tabor, Tad Lincoln, Takarazuka Family Land, Tassedethe, Tastesoon, The High Fin Sperm Whale, The wub, TheHighTree, TheYmode, ThreeE, Throwaway85, Thuglas, Timrem, Tmaull, Tommy2010, Trelvis, Twredfish, UNV, Vanished user 47736712, WLU, WhatamIdoing, Whkoh, WillBecker, WvEngen, Zero1328, ^demon, 274 anonymous edits Magnetic dipoledipole interaction Source: http://en.wikipedia.org/w/index.php?oldid=467027238 Contributors: Barticus88, Leo C Stein, Macholl, Sangak, Scray, Warbler271, 15 anonymous edits Magnetic hyperthermia Source: http://en.wikipedia.org/w/index.php?oldid=488108689 Contributors: AnnaJune, Arcadian, Auntof6, Gaius Cornelius, LilHelpa, Mandarax, Marvinandmilo, Pdreiss, Rich Farmbrough, RockMagnetist, Roulianne, WhatamIdoing, 18 anonymous edits Magnetic nanoparticles Source: http://en.wikipedia.org/w/index.php?oldid=489167348 Contributors: Antony-22, Beautygab, Biscuittin, Bunnyhop11, CKarnstein, Carlog3, CommonsDelinker, Craigsjones, EoGuy, GoingBatty, Igoon, Jbonastre, Jojalozzo, Kcoopersmith, Kunnskap, Lamro, Materialscientist, Nuriapar, Paulbrock, Philip john123, Rjwilmsi, Roulianne, Sgoofy8, Supermaster2011, WBardwin, WereSpielChequers, Whosasking, Wikieditonator, Woohookitty, Yaser hassan 2006, Yobmod, 18 anonymous edits Single-molecule magnet Source: http://en.wikipedia.org/w/index.php?oldid=480924578 Contributors: Awickert, Bloomy717, Buzz-tardis, Christian75, Cimorcus, Daisystanton, Dougszathkey, DragonflySixtyseven, Dreamer08, Embec, Enric Naval, Giraffedata, Gobonobo, Headbomb, Heron, Hollyev, Intgr, Lantonov, MagnInd, Materialscientist, Mereda, Nickptar, Rayc, Rbrausse, Reedijk, RockMagnetist, Takaaki, Thatjenn, Tiglet, Tomgally, V8rik, Venny85, WAS 4.250, 45 anonymous edits Magnetic anisotropy Source: http://en.wikipedia.org/w/index.php?oldid=487155027 Contributors: Amtiss, Antariki Vandanamu, Bearcat, CrusCuspis, Dondegroovily, Fbprickett, GregorB, Hbent, LokiClock, Marechad, Nick Number, Nick84, RockMagnetist, Tamtamar, Verkhovensky, WvEngen, Zhangzhe0101, 6 anonymous edits Magnetocrystalline anisotropy Source: http://en.wikipedia.org/w/index.php?oldid=463507203 Contributors: Courcelles, Dondegroovily, Icairns, Jag123, Nick Number, Numbah1, RockMagnetist, Salsb, Schmiteye, ShuoCheng, Tone, Vibhav Chauhan, W2023, WvEngen, Yangtzushun, 12 anonymous edits Electron configuration Source: http://en.wikipedia.org/w/index.php?oldid=489533508 Contributors: 2over0, 4twenty42o, AManWithNoPlan, Ace of Spades IV, Adventurer, Aeluwas, Af648, Aitias, Aldaron, Aleksa Lukic, Alex43223, Alexjohnc3, AlphaEta, Andre Engels, Andres, Ankit.sunrise, Anna Lincoln, Anoop.m, Antandrus, Anthony Appleyard, Anville, Appy3, Arag0rn, Aritate, Arthur Rubin, Ashley Y, Astronautics, AtheWeatherman, Aushulz, Azure8472, Baccyak4H, Bduke, Beetstra, Bender235, Bhludzin, Bliz1, Bomac, Brian0918, Bth, Bucephalus, Byrgenwulf, Calmargulis, CardinalDan, Cburnett, Cerebellum, CharlotteWebb, Chau7, Chris the speller, Chuckiesdad, Cinderblock63, Coffee, Colin Kimbrell, Crazycomputers, Cremepuff222, Cstaffa, DARTH SIDIOUS 2, DMacks, Dan653, Dark Shikari, Darkwind, DavidCBryant, Davidwhite18, Ddoherty, Deagle AP, Diannaa, Dipanshu.sheru, Dirac1933, Dirac66, Discospinster, Djdaedalus, Djnjwd, Djr32, Dnvrfantj, Donarreiskoffer, Donsimon, Double sharp, DrJolo, Dreg743, Drova, Dumoren, Dysprosia, EEMIV, EagleFan, Earlypsychosis, Epbr123, Erik Zachte, FT2, Falcon8765, Femto, Footballfan190, Fresheneesz, Fucktosh, Fvw, Fzhi555, Gail, GaryW, Gentgeen, Giftlite, Glane23, Glenn L, GrahamColm, Hadal, HansHermans, Headbomb, Holyhell5050, Hugo-cs, IdealEric, Ilyushka88, Iridescent, J.delanoy, Jacek Kendysz, Janek Kozicki, Jayko, Jelson25, Jengelh, John, John Reaves, Johndburger, JohnofPhoenix, Juliancolton, KSmrq, Karada, KarlHegbloom, KeepItClean, Kevkevkevkev, Kku, Kmarinas86, Kouhoutek, Krazywrath, Kungfuadam, Kurzon, Kwamikagami, LMB, Lanthanum-138, Linas, Lir, LokiClock, Looxix, Lucky 6.9, Luna Santin, MJ94, MZMcBride, Mako098765, Malbi, Manifolded, Marik7772003, Materialscientist, Matt18224, Mattopia, Mboverload, Meddlingwithfire, Michael Hardy, Mor, Mpatel, MrChupon, Mrh30, Mttcmbs, Nakon, Nantaskot, NawlinWiki, Nayuki, Necessary Evil, Nk, Nrcprm2026, Nsmith 84, NuclearWarfare, OMCV, Ojigiri, Oxymoron83, P. B. Mann, Patrick, Patstuart, PhilKnight, Physchim62, Pinkadelica, Pjstewart, Prari, Putodog, Qbmaster, Quibik, Qwyrxian, R3m0t, RG2, RJaguar3, RMFan1, RadioFan, Radon210, RainbowOfLight, Rdsmith4, Redrose64, Res2216firestar, Retireduser1111, RexNL, Robth, Romanm, Rrburke, S h i v a (Visnu), Samanthaclark11, Sbharris, Schmloof, Schwnj, Seaphoto, Sergio.ballestrero, Shadowjams, SimonP, Skizzik, Smack, SmilesALot, Snoyes, Spacecadet262, Spoon!, Steviedpeele, StradivariusTV, Svick, TBadger, Tamariandre, Teles, Tesi1700, Tetracube, The Anome, The High Fin Sperm Whale, The Thing That Should Not Be, Tide rolls, Timemutt, Timwi, Tiptoety, Tobias Bergemann, TokioHotel93, Topbanana, Trevorhailey1, Tyco.skinner, Ulric1313, Unfree, V8rik, Vb, Vsmith, Wavelength, Weekwhom, WhiteTimberwolf, Wikipelli, X42bn6, Yannledu, Yelyos, Yuk ngan, , , , , 557 anonymous edits Inverse magnetostrictive effect Source: http://en.wikipedia.org/w/index.php?oldid=452950596 Contributors: Kradak, Marechad, Melchoir, RockMagnetist, 1 anonymous edits Exchange bias Source: http://en.wikipedia.org/w/index.php?oldid=452947787 Contributors: Chaiken, Gredig, Hooperbloob, John, LarRan, Leowiz, Liquidcentre, Porqin, RockMagnetist, Schmiteye, WvEngen, 8 anonymous edits

256

Image Sources, Licenses and Contributors

257

Image Sources, Licenses and Contributors


File:M Faraday Th Phillips oil 1842.jpg Source: http://en.wikipedia.org/w/index.php?title=File:M_Faraday_Th_Phillips_oil_1842.jpg License: Public Domain Contributors: Thomas Phillips File:Magnetism.svg Source: http://en.wikipedia.org/w/index.php?title=File:Magnetism.svg License: Public Domain Contributors: User:John Aplessed File:Ferromagneses penzermek 1.jpg Source: http://en.wikipedia.org/w/index.php?title=File:Ferromagneses_penzermek_1.jpg License: Creative Commons Attribution-Sharealike 3.0 Contributors: Ztonyi Sndor, (ifj.) Fizped Image:Ferromag Matl Sketch.JPG Source: http://en.wikipedia.org/w/index.php?title=File:Ferromag_Matl_Sketch.JPG License: Creative Commons Attribution 2.5 Contributors: Original uploader was JA.Davidson at en.wikipedia Image:Ferromag Matl Magnetized.JPG Source: http://en.wikipedia.org/w/index.php?title=File:Ferromag_Matl_Magnetized.JPG License: Creative Commons Attribution 2.5 Contributors: Original uploader was JA.Davidson at en.wikipedia Image:Antiferromagnetic ordering.svg Source: http://en.wikipedia.org/w/index.php?title=File:Antiferromagnetic_ordering.svg License: Creative Commons Attribution-ShareAlike 3.0 Unported Contributors: Michael Schmid Image:Ferrimagnetic ordering.svg Source: http://en.wikipedia.org/w/index.php?title=File:Ferrimagnetic_ordering.svg License: Creative Commons Attribution-ShareAlike 3.0 Unported Contributors: Michael Schmid File:Electromagnet.gif Source: http://en.wikipedia.org/w/index.php?title=File:Electromagnet.gif License: unknown Contributors: Anynobody Image:Magnet0873.png Source: http://en.wikipedia.org/w/index.php?title=File:Magnet0873.png License: Public Domain Contributors: Newton Henry Black Image:Descartes magnetic field.jpg Source: http://en.wikipedia.org/w/index.php?title=File:Descartes_magnetic_field.jpg License: Public Domain Contributors: Ren Descartes Image:Magnetic field near pole.svg Source: http://en.wikipedia.org/w/index.php?title=File:Magnetic_field_near_pole.svg License: Creative Commons Attribution-Sharealike 3.0 Contributors: TStein Image:VFPt dipole electric.svg Source: http://en.wikipedia.org/w/index.php?title=File:VFPt_dipole_electric.svg License: Creative Commons Attribution-Sharealike 3.0 Contributors: Geek3 Image:VFPt dipole magnetic3.svg Source: http://en.wikipedia.org/w/index.php?title=File:VFPt_dipole_magnetic3.svg License: Creative Commons Attribution-Sharealike 3.0 Contributors: Geek3 image:Dipole in uniform H field.svg Source: http://en.wikipedia.org/w/index.php?title=File:Dipole_in_uniform_H_field.svg License: Creative Commons Zero Contributors: User:Fred the Oyster Image:cross parallelogram.png Source: http://en.wikipedia.org/w/index.php?title=File:Cross_parallelogram.png License: Public Domain Contributors: Oleg Alexandrov Image:Manoderecha.svg Source: http://en.wikipedia.org/w/index.php?title=File:Manoderecha.svg License: GNU Free Documentation License Contributors: Jfmelero Image:Solenoid-1 (vertical).png Source: http://en.wikipedia.org/w/index.php?title=File:Solenoid-1_(vertical).png License: Public Domain Contributors: Zureks File:Magnetic force.svg Source: http://en.wikipedia.org/w/index.php?title=File:Magnetic_force.svg License: Creative Commons Attribution-Sharealike 3.0 Contributors: User:Maschen Image:Regla mano derecha Laplace.svg Source: http://en.wikipedia.org/w/index.php?title=File:Regla_mano_derecha_Laplace.svg License: Creative Commons Attribution-Sharealike 3.0 Contributors: Jfmelero File:BIsAPseudovector.svg Source: http://en.wikipedia.org/w/index.php?title=File:BIsAPseudovector.svg License: Public Domain Contributors: Sbyrnes321 Image:Earths Magnetic Field Confusion.svg Source: http://en.wikipedia.org/w/index.php?title=File:Earths_Magnetic_Field_Confusion.svg License: Creative Commons Attribution-Sharealike 3.0 Contributors: TStein Image:Magnetic quadrupole moment.svg Source: http://en.wikipedia.org/w/index.php?title=File:Magnetic_quadrupole_moment.svg License: Public Domain Contributors: Original uploader was K. Aainsqatsi at en.wikipedia Image:Magnetic moment.svg Source: http://en.wikipedia.org/w/index.php?title=File:Magnetic_moment.svg License: Creative Commons Attribution-Sharealike 3.0 Contributors: User:Glosser.ca Image:Solenoid, air core, insulated, 20 turns, (shaded).svg Source: http://en.wikipedia.org/w/index.php?title=File:Solenoid,_air_core,_insulated,_20_turns,_(shaded).svg License: Public Domain Contributors: Inductiveload Image:Magnetic ring dipole field lines.svg Source: http://en.wikipedia.org/w/index.php?title=File:Magnetic_ring_dipole_field_lines.svg License: Creative Commons Attribution-Sharealike 3.0,2.5,2.0,1.0 Contributors: Sharayanan Image:Dipoledipole.svg Source: http://en.wikipedia.org/w/index.php?title=File:Dipoledipole.svg License: Public Domain Contributors: Kakila File:SingleDomainMagneticCharges.svg Source: http://en.wikipedia.org/w/index.php?title=File:SingleDomainMagneticCharges.svg License: Creative Commons Zero Contributors: RockMagnetist Image:ClosureDomainDemo.svg Source: http://en.wikipedia.org/w/index.php?title=File:ClosureDomainDemo.svg License: Creative Commons Zero Contributors: RockMagnetist File:Permeability by Zureks.svg Source: http://en.wikipedia.org/w/index.php?title=File:Permeability_by_Zureks.svg License: Public Domain Contributors: Zureks File:Permeability of ferromagnet by Zureks.svg Source: http://en.wikipedia.org/w/index.php?title=File:Permeability_of_ferromagnet_by_Zureks.svg License: Public Domain Contributors: Zureks Image:VFPt cylindrical magnet thumb.svg Source: http://en.wikipedia.org/w/index.php?title=File:VFPt_cylindrical_magnet_thumb.svg License: Creative Commons Attribution-Sharealike 3.0 Contributors: Geek3 Image:VFPt cylindrical magnets attracting.svg Source: http://en.wikipedia.org/w/index.php?title=File:VFPt_cylindrical_magnets_attracting.svg License: Creative Commons Attribution-Sharealike 3.0 Contributors: Geek3 Image:VFPt cylindrical magnets repelling.svg Source: http://en.wikipedia.org/w/index.php?title=File:VFPt_cylindrical_magnets_repelling.svg License: Creative Commons Attribution-Sharealike 3.0 Contributors: Geek3 Image:VFPt dipole thumb.svg Source: http://en.wikipedia.org/w/index.php?title=File:VFPt_dipole_thumb.svg License: Creative Commons Attribution-Sharealike 3.0 Contributors: Geek3 File:Lorentz force.svg Source: http://en.wikipedia.org/w/index.php?title=File:Lorentz_force.svg License: GNU Free Documentation License Contributors: User:Jaro.p File:Cyclotron motion.jpg Source: http://en.wikipedia.org/w/index.php?title=File:Cyclotron_motion.jpg License: Creative Commons Attribution-Share Alike Contributors: Marcin Biaek File:Regla mano derecha Laplace.svg Source: http://en.wikipedia.org/w/index.php?title=File:Regla_mano_derecha_Laplace.svg License: Creative Commons Attribution-Sharealike 3.0 Contributors: Jfmelero File:Gouy bal.png Source: http://en.wikipedia.org/w/index.php?title=File:Gouy_bal.png License: Creative Commons Zero Contributors: Petergans (talk) File:Susceptibility.png Source: http://en.wikipedia.org/w/index.php?title=File:Susceptibility.png License: Creative Commons Zero Contributors: Petergans (talk) File:Copper(II)-acetate-3D-balls2.png Source: http://en.wikipedia.org/w/index.php?title=File:Copper(II)-acetate-3D-balls2.png License: Public Domain Contributors: Copper(II)-acetate-3D-balls.png: Benjah-bmm27 derivative work: Materialscientist (talk) File:Ferrimagnetic ordering.svg Source: http://en.wikipedia.org/w/index.php?title=File:Ferrimagnetic_ordering.svg License: Creative Commons Attribution-ShareAlike 3.0 Unported Contributors: Michael Schmid File:Antiferromagnetic ordering.svg Source: http://en.wikipedia.org/w/index.php?title=File:Antiferromagnetic_ordering.svg License: Creative Commons Attribution-ShareAlike 3.0 Unported Contributors: Michael Schmid File:CFT - Low Spin Splitting Diagram 2.png Source: http://en.wikipedia.org/w/index.php?title=File:CFT_-_Low_Spin_Splitting_Diagram_2.png License: GNU Free Documentation License Contributors: YanA File:CFT - High Spin Splitting Diagram 2.png Source: http://en.wikipedia.org/w/index.php?title=File:CFT_-_High_Spin_Splitting_Diagram_2.png License: GNU Free Documentation License Contributors: YanA File:EPR methyl.png Source: http://en.wikipedia.org/w/index.php?title=File:EPR_methyl.png License: Public Domain Contributors: EPR_methyl.jpg: Astrochemist derivative work: Mikhail Ryazanov (talk) File:MTSL chemical structure.png Source: http://en.wikipedia.org/w/index.php?title=File:MTSL_chemical_structure.png License: Public Domain Contributors: Cacycle Image:Equilibrium.svg Source: http://en.wikipedia.org/w/index.php?title=File:Equilibrium.svg License: Public Domain Contributors: L'Aquatique

Image Sources, Licenses and Contributors


Image:Periodic Table with unpaired electrons.svg Source: http://en.wikipedia.org/w/index.php?title=File:Periodic_Table_with_unpaired_electrons.svg License: Public Domain Contributors: KES47 Image:neon orbitals.JPG Source: http://en.wikipedia.org/w/index.php?title=File:Neon_orbitals.JPG License: Public Domain Contributors: Benjah-bmm27, Mortadelo2005, 1 anonymous edits Image:Bohr-atom-PAR.svg Source: http://en.wikipedia.org/w/index.php?title=File:Bohr-atom-PAR.svg License: unknown Contributors: Original uploader was JabberWok at en.wikipedia File:HydrogenOrbitalsN6L0M0.png Source: http://en.wikipedia.org/w/index.php?title=File:HydrogenOrbitalsN6L0M0.png License: GNU General Public License Contributors: Carlo Barraco File:D orbitals.svg Source: http://en.wikipedia.org/w/index.php?title=File:D_orbitals.svg License: Creative Commons Attribution-ShareAlike 3.0 Unported Contributors: Benjah-bmm27, Matthias M., Pieter Kuiper, Sven, 1 anonymous edits Image:S1M0.png Source: http://en.wikipedia.org/w/index.php?title=File:S1M0.png License: Public Domain Contributors: Dhatfield Image:S2M0.png Source: http://en.wikipedia.org/w/index.php?title=File:S2M0.png License: Public Domain Contributors: Dhatfield Image:P2M0.png Source: http://en.wikipedia.org/w/index.php?title=File:P2M0.png License: Public Domain Contributors: Dhatfield Image:P2M1.png Source: http://en.wikipedia.org/w/index.php?title=File:P2M1.png License: Public Domain Contributors: Dhatfield Image:P2M-1.png Source: http://en.wikipedia.org/w/index.php?title=File:P2M-1.png License: Public Domain Contributors: Dhatfield Image:S3M0.png Source: http://en.wikipedia.org/w/index.php?title=File:S3M0.png License: Public Domain Contributors: Dhatfield Image:P3M0.png Source: http://en.wikipedia.org/w/index.php?title=File:P3M0.png License: Public Domain Contributors: Dhatfield Image:P3M1.png Source: http://en.wikipedia.org/w/index.php?title=File:P3M1.png License: Public Domain Contributors: Dhatfield Image:P3M-1.png Source: http://en.wikipedia.org/w/index.php?title=File:P3M-1.png License: Public Domain Contributors: Dhatfield Image:D3M0.png Source: http://en.wikipedia.org/w/index.php?title=File:D3M0.png License: Public Domain Contributors: Dhatfield Image:D3M1.png Source: http://en.wikipedia.org/w/index.php?title=File:D3M1.png License: Public Domain Contributors: Dhatfield Image:D3M-1.png Source: http://en.wikipedia.org/w/index.php?title=File:D3M-1.png License: Public Domain Contributors: Dhatfield Image:D3M2.png Source: http://en.wikipedia.org/w/index.php?title=File:D3M2.png License: Public Domain Contributors: Dhatfield Image:D3M-2.png Source: http://en.wikipedia.org/w/index.php?title=File:D3M-2.png License: Public Domain Contributors: Dhatfield Image:S4M0.png Source: http://en.wikipedia.org/w/index.php?title=File:S4M0.png License: Public Domain Contributors: Dhatfield Image:P4M0.png Source: http://en.wikipedia.org/w/index.php?title=File:P4M0.png License: Public Domain Contributors: Dhatfield Image:P4M1.png Source: http://en.wikipedia.org/w/index.php?title=File:P4M1.png License: Public Domain Contributors: Dhatfield, Pieter Kuiper Image:P4M-1.png Source: http://en.wikipedia.org/w/index.php?title=File:P4M-1.png License: Public Domain Contributors: Dhatfield Image:D4M0.png Source: http://en.wikipedia.org/w/index.php?title=File:D4M0.png License: Public Domain Contributors: Dhatfield Image:D4M1.png Source: http://en.wikipedia.org/w/index.php?title=File:D4M1.png License: Public Domain Contributors: Dhatfield Image:D4M-1.png Source: http://en.wikipedia.org/w/index.php?title=File:D4M-1.png License: Public Domain Contributors: Dhatfield Image:D4M2.png Source: http://en.wikipedia.org/w/index.php?title=File:D4M2.png License: Public Domain Contributors: Dhatfield Image:D4M-2.png Source: http://en.wikipedia.org/w/index.php?title=File:D4M-2.png License: Public Domain Contributors: Dhatfield Image:F4M0.png Source: http://en.wikipedia.org/w/index.php?title=File:F4M0.png License: Public Domain Contributors: Dhatfield Image:F4M1.png Source: http://en.wikipedia.org/w/index.php?title=File:F4M1.png License: Public Domain Contributors: Dhatfield Image:F4M-1.png Source: http://en.wikipedia.org/w/index.php?title=File:F4M-1.png License: Public Domain Contributors: Dhatfield Image:F4M2.png Source: http://en.wikipedia.org/w/index.php?title=File:F4M2.png License: Public Domain Contributors: Dhatfield Image:F4M-2.png Source: http://en.wikipedia.org/w/index.php?title=File:F4M-2.png License: Public Domain Contributors: Dhatfield Image:F4M3.png Source: http://en.wikipedia.org/w/index.php?title=File:F4M3.png License: Public Domain Contributors: Dhatfield Image:F4M-3.png Source: http://en.wikipedia.org/w/index.php?title=File:F4M-3.png License: Public Domain Contributors: Dhatfield Image:S5M0.png Source: http://en.wikipedia.org/w/index.php?title=File:S5M0.png License: Public Domain Contributors: Dhatfield Image:P5M0.png Source: http://en.wikipedia.org/w/index.php?title=File:P5M0.png License: Public Domain Contributors: Dhatfield Image:P5M1.png Source: http://en.wikipedia.org/w/index.php?title=File:P5M1.png License: Public Domain Contributors: Dhatfield Image:P5M-1.png Source: http://en.wikipedia.org/w/index.php?title=File:P5M-1.png License: Public Domain Contributors: Dhatfield Image:D5M0.png Source: http://en.wikipedia.org/w/index.php?title=File:D5M0.png License: Public Domain Contributors: Dhatfield Image:D5M1.png Source: http://en.wikipedia.org/w/index.php?title=File:D5M1.png License: Public Domain Contributors: Dhatfield Image:D5M-1.png Source: http://en.wikipedia.org/w/index.php?title=File:D5M-1.png License: Public Domain Contributors: Dhatfield Image:D5M2.png Source: http://en.wikipedia.org/w/index.php?title=File:D5M2.png License: Public Domain Contributors: Dhatfield Image:D5M-2.png Source: http://en.wikipedia.org/w/index.php?title=File:D5M-2.png License: Public Domain Contributors: Dhatfield Image:S6M0.png Source: http://en.wikipedia.org/w/index.php?title=File:S6M0.png License: Public Domain Contributors: Dhatfield Image:P6M0.png Source: http://en.wikipedia.org/w/index.php?title=File:P6M0.png License: Public Domain Contributors: Dhatfield Image:P6M1.png Source: http://en.wikipedia.org/w/index.php?title=File:P6M1.png License: Public Domain Contributors: Dhatfield Image:P6M-1.png Source: http://en.wikipedia.org/w/index.php?title=File:P6M-1.png License: Public Domain Contributors: Dhatfield Image:S7M0.png Source: http://en.wikipedia.org/w/index.php?title=File:S7M0.png License: Public Domain Contributors: Dhatfield Image:Drum vibration mode01.gif Source: http://en.wikipedia.org/w/index.php?title=File:Drum_vibration_mode01.gif License: Public Domain Contributors: Oleg Alexandrov Image:Drum vibration mode02.gif Source: http://en.wikipedia.org/w/index.php?title=File:Drum_vibration_mode02.gif License: Public Domain Contributors: Oleg Alexandrov Image:Drum vibration mode03.gif Source: http://en.wikipedia.org/w/index.php?title=File:Drum_vibration_mode03.gif License: Public Domain Contributors: Oleg Alexandrov Image:Drum vibration mode11.gif Source: http://en.wikipedia.org/w/index.php?title=File:Drum_vibration_mode11.gif License: Public Domain Contributors: Oleg Alexandrov Image:Drum vibration mode12.gif Source: http://en.wikipedia.org/w/index.php?title=File:Drum_vibration_mode12.gif License: Public Domain Contributors: Oleg Alexandrov Image:Drum vibration mode13.gif Source: http://en.wikipedia.org/w/index.php?title=File:Drum_vibration_mode13.gif License: Public Domain Contributors: Oleg Alexandrov Image:Drum vibration mode21.gif Source: http://en.wikipedia.org/w/index.php?title=File:Drum_vibration_mode21.gif License: Public Domain Contributors: Oleg Alexandrov Image:Drum vibration mode22.gif Source: http://en.wikipedia.org/w/index.php?title=File:Drum_vibration_mode22.gif License: Public Domain Contributors: Oleg Alexandrov Image:Drum vibration mode23.gif Source: http://en.wikipedia.org/w/index.php?title=File:Drum_vibration_mode23.gif License: Public Domain Contributors: Oleg Alexandrov Image:Electron orbitals.svg Source: http://en.wikipedia.org/w/index.php?title=File:Electron_orbitals.svg License: Public Domain Contributors: Patricia.fidi Image:LFTi(III).png Source: http://en.wikipedia.org/w/index.php?title=File:LFTi(III).png License: Public Domain Contributors: Original uploader was Smokefoot at en.wikipedia Image:CFT - High Spin Splitting Diagram 2.png Source: http://en.wikipedia.org/w/index.php?title=File:CFT_-_High_Spin_Splitting_Diagram_2.png License: GNU Free Documentation License Contributors: YanA File:Klechkovski rule.svg Source: http://en.wikipedia.org/w/index.php?title=File:Klechkovski_rule.svg License: Creative Commons Attribution-ShareAlike 3.0 Unported Contributors: Bono, Kilom691, Pieter Kuiper Image:Sommerfeld ellipses.svg Source: http://en.wikipedia.org/w/index.php?title=File:Sommerfeld_ellipses.svg License: Public Domain Contributors: Pieter Kuiper File:Cisplatin-3D-balls.png Source: http://en.wikipedia.org/w/index.php?title=File:Cisplatin-3D-balls.png License: Public Domain Contributors: Benjah-bmm27 File:Hexol-2D-wedged.png Source: http://en.wikipedia.org/w/index.php?title=File:Hexol-2D-wedged.png License: Public Domain Contributors: Ben Mills Image:Cis-dichlorotetraamminecobalt(III).png Source: http://en.wikipedia.org/w/index.php?title=File:Cis-dichlorotetraamminecobalt(III).png License: Public Domain Contributors: Benjah-bmm27

258

Image Sources, Licenses and Contributors


Image:Trans-dichlorotetraamminecobalt(III).png Source: http://en.wikipedia.org/w/index.php?title=File:Trans-dichlorotetraamminecobalt(III).png License: Public Domain Contributors: Benjah-bmm27 Image:Fac-trichlorotriamminecobalt(III).png Source: http://en.wikipedia.org/w/index.php?title=File:Fac-trichlorotriamminecobalt(III).png License: Public Domain Contributors: Benjah-bmm27 Image:Mer-trichlorotriamminecobalt(III).png Source: http://en.wikipedia.org/w/index.php?title=File:Mer-trichlorotriamminecobalt(III).png License: Public Domain Contributors: Benjah-bmm27 Image:Delta-tris(oxalato)ferrate(III)-3D-balls.png Source: http://en.wikipedia.org/w/index.php?title=File:Delta-tris(oxalato)ferrate(III)-3D-balls.png License: Public Domain Contributors: Benjah-bmm27, Jynto, Tpa2067, Wickey-nl Image:Lambda-tris(oxalato)ferrate(III)-3D-balls.png Source: http://en.wikipedia.org/w/index.php?title=File:Lambda-tris(oxalato)ferrate(III)-3D-balls.png License: Public Domain Contributors: Benjah-bmm27, Jynto, Wickey-nl Image:Delta-cis-dichlorobis(ethylenediamine)cobalt(III).png Source: http://en.wikipedia.org/w/index.php?title=File:Delta-cis-dichlorobis(ethylenediamine)cobalt(III).png License: Public Domain Contributors: Benjah-bmm27 Image:Lambda-cis-dichlorobis(ethylenediamine)cobalt(III).png Source: http://en.wikipedia.org/w/index.php?title=File:Lambda-cis-dichlorobis(ethylenediamine)cobalt(III).png License: Public Domain Contributors: Benjah-bmm27 Image:Copper complex.jpg Source: http://en.wikipedia.org/w/index.php?title=File:Copper_complex.jpg License: Public Domain Contributors: Xiaopo Image:Diamagnetic graphite levitation.jpg Source: http://en.wikipedia.org/w/index.php?title=File:Diamagnetic_graphite_levitation.jpg License: Public domain Contributors: en:User:Splarka Image:Superconductor.GIF Source: http://en.wikipedia.org/w/index.php?title=File:Superconductor.GIF License: Public Domain Contributors: David Meeker wrote FEMM 4.2 Image:Frog diamagnetic levitation.jpg Source: http://en.wikipedia.org/w/index.php?title=File:Frog_diamagnetic_levitation.jpg License: GNU Free Documentation License Contributors: Lijnis Nelemans Image:Paramagnetic probe without magnetic field.svg Source: http://en.wikipedia.org/w/index.php?title=File:Paramagnetic_probe_without_magnetic_field.svg License: Public Domain Contributors: Jens Bning (Jensel) File:Paramagnetism of liquid oxygen.jpeg Source: http://en.wikipedia.org/w/index.php?title=File:Paramagnetism_of_liquid_oxygen.jpeg License: Public Domain Contributors: Pieter Kuiper Image:Para-ferro-anti.jpg Source: http://en.wikipedia.org/w/index.php?title=File:Para-ferro-anti.jpg License: GNU Free Documentation License Contributors: NlJcwf Image:magnetization2.jpg Source: http://en.wikipedia.org/w/index.php?title=File:Magnetization2.jpg License: Public Domain Contributors: Zedh Image:Langevin function.png Source: http://en.wikipedia.org/w/index.php?title=File:Langevin_function.png License: Creative Commons Attribution-Sharealike 3.0 Contributors: Ddcampayo at en.wikipedia Image:EPR spectometer.JPG Source: http://en.wikipedia.org/w/index.php?title=File:EPR_spectometer.JPG License: GNU Free Documentation License Contributors: Photo made by Przemyslaw "Tukan" Grudnik Image:EPR splitting.jpg Source: http://en.wikipedia.org/w/index.php?title=File:EPR_splitting.jpg License: Public Domain Contributors: Original uploader was Astrochemist at en.wikipedia Image:EPR lines.png Source: http://en.wikipedia.org/w/index.php?title=File:EPR_lines.png License: Public Domain Contributors: EPR_lines.jpg: Original uploader was Astrochemist at en.wikipedia derivative work: Mikhail Ryazanov (talk) Image:EPR methyl.png Source: http://en.wikipedia.org/w/index.php?title=File:EPR_methyl.png License: Public Domain Contributors: EPR_methyl.jpg: Astrochemist derivative work: Mikhail Ryazanov (talk) Image:EPR methoxymethyl.png Source: http://en.wikipedia.org/w/index.php?title=File:EPR_methoxymethyl.png License: Public Domain Contributors: Original JPEG: http://en.wikipedia.org/wiki/User:Astrochemist Image:EPR multifrequency spectra.png Source: http://en.wikipedia.org/w/index.php?title=File:EPR_multifrequency_spectra.png License: Public Domain Contributors: Original JPEG: http://en.wikipedia.org/wiki/User:Astrochemist File:Przession2.png Source: http://en.wikipedia.org/w/index.php?title=File:Przession2.png License: Public Domain Contributors: Yamavu (talk) Image:HahnEcho GWM.gif Source: http://en.wikipedia.org/w/index.php?title=File:HahnEcho_GWM.gif License: Creative Commons Attribution-Sharealike 3.0 Contributors: GavinMorley File:Spin-physics-w.svg Source: http://en.wikipedia.org/w/index.php?title=File:Spin-physics-w.svg License: Public Domain Contributors: Kes47 (?) File:"Father" and "Mother" of the series "Spin Family".jpg Source: http://en.wikipedia.org/w/index.php?title=File:"Father"_and_"Mother"_of_the_series_"Spin_Family".jpg License: Creative Commons Attribution-Sharealike 3.0 Contributors: User:Julianva File:Wolfgang Pauli young.jpg Source: http://en.wikipedia.org/w/index.php?title=File:Wolfgang_Pauli_young.jpg License: Public Domain Contributors: Dodo, Ephraim33, Eusebius, Harp, JdH Image:The muon g-2.svg Source: http://en.wikipedia.org/w/index.php?title=File:The_muon_g-2.svg License: Creative Commons Attribution 3.0 Contributors: User A1 at en.wikipedia File:Breit-rabi-Zeeman.png Source: http://en.wikipedia.org/w/index.php?title=File:Breit-rabi-Zeeman.png License: Creative Commons Attribution-Sharealike 3.0 Contributors: User:Danski14 File:Loudspeaker.svg Source: http://en.wikipedia.org/w/index.php?title=File:Loudspeaker.svg License: Public Domain Contributors: Bayo, Gmaxwell, Husky, Iamunknown, Mirithing, Myself488, Nethac DIU, Omegatron, Rocket000, The Evil IP address, Wouterhagens, 18 anonymous edits Image:Zeeman p s doublet.svg Source: http://en.wikipedia.org/w/index.php?title=File:Zeeman_p_s_doublet.svg License: Creative Commons Attribution 2.5 Contributors: Original uploader was Evgeny at en.wikipedia File:Sunzeeman1919.png Source: http://en.wikipedia.org/w/index.php?title=File:Sunzeeman1919.png License: Public Domain Contributors: Hale File:Solar storm 2003-10-28 (SOHO-MDI, Magnetograms).png Source: http://en.wikipedia.org/w/index.php?title=File:Solar_storm_2003-10-28_(SOHO-MDI,_Magnetograms).png License: Public Domain Contributors: NASA/Goddard Space Flight Center Scientific Visualization Studio File:HAtomOrbitals.png Source: http://en.wikipedia.org/w/index.php?title=File:HAtomOrbitals.png License: GNU Free Documentation License Contributors: Admrboltz, Benjah-bmm27, Dbc334, Dbenbenn, Ejdzej, Falcorian, Kborland, MichaelDiederich, Mion, Saperaud, 6 anonymous edits File:Circular_Standing_Wave.gif Source: http://en.wikipedia.org/w/index.php?title=File:Circular_Standing_Wave.gif License: Creative Commons Attribution-ShareAlike 3.0 Unported Contributors: Yuta Aoki File:Vector model of orbital angular momentum.svg Source: http://en.wikipedia.org/w/index.php?title=File:Vector_model_of_orbital_angular_momentum.svg License: Creative Commons Attribution-Sharealike 3.0 Contributors: User:Maschen File:RotationOperators.svg Source: http://en.wikipedia.org/w/index.php?title=File:RotationOperators.svg License: Creative Commons Zero Contributors: User:Sbyrnes321 File:Gyroskop.jpg Source: http://en.wikipedia.org/w/index.php?title=File:Gyroskop.jpg License: GNU Free Documentation License Contributors: Kiko2000, Opponent, Tano4595, WikipediaMaster File:Torque animation.gif Source: http://en.wikipedia.org/w/index.php?title=File:Torque_animation.gif License: Public Domain Contributors: Yawe File:Angular momentum definition.svg Source: http://en.wikipedia.org/w/index.php?title=File:Angular_momentum_definition.svg License: Creative Commons Attribution-Sharealike 3.0 Contributors: User:Maschen Image:Cup of Russia 2010 - Yuko Kawaguti (2).jpg Source: http://en.wikipedia.org/w/index.php?title=File:Cup_of_Russia_2010_-_Yuko_Kawaguti_(2).jpg License: Creative Commons Zero Contributors: deerstop File:Angular momentum conservation.svg Source: http://en.wikipedia.org/w/index.php?title=File:Angular_momentum_conservation.svg License: Creative Commons Attribution-Sharealike 3.0 Contributors: User:Maschen File:PrecessionOfATop.svg Source: http://en.wikipedia.org/w/index.php?title=File:PrecessionOfATop.svg License: Creative Commons Attribution-Sharealike 2.5 Contributors: Xavier Snelgrove File:Circular Standing Wave.gif Source: http://en.wikipedia.org/w/index.php?title=File:Circular_Standing_Wave.gif License: Creative Commons Attribution-ShareAlike 3.0 Unported Contributors: Yuta Aoki Image:Ferrofluid Magnet under glass edit.jpg Source: http://en.wikipedia.org/w/index.php?title=File:Ferrofluid_Magnet_under_glass_edit.jpg License: GNU Free Documentation License Contributors: Gregory F. Maxwell < gmaxwell@gmail.com> File:Ferrofluid poles.jpg Source: http://en.wikipedia.org/w/index.php?title=File:Ferrofluid_poles.jpg License: GNU Free Documentation License Contributors: Gregory F. Maxwell < gmaxwell@gmail.com>

259

Image Sources, Licenses and Contributors


Image:Ferrofluid close.jpg Source: http://en.wikipedia.org/w/index.php?title=File:Ferrofluid_close.jpg License: Creative Commons Attribution 3.0 Contributors: Opoterser Image:Ferrofluid in magnetic field.jpg Source: http://en.wikipedia.org/w/index.php?title=File:Ferrofluid_in_magnetic_field.jpg License: Creative Commons Attribution 2.0 Contributors: Steve Jurvetson Image:Cobalt-graphene-nanoparticle.tif Source: http://en.wikipedia.org/w/index.php?title=File:Cobalt-graphene-nanoparticle.tif License: Creative Commons Attribution-Sharealike 3.0 Contributors: Supermaster2011 Image:Flame-spray-synthesis.JPG Source: http://en.wikipedia.org/w/index.php?title=File:Flame-spray-synthesis.JPG License: Creative Commons Zero Contributors: Evagelos Athanassiou Image:Flame-spray-synthesis-2.JPG Source: http://en.wikipedia.org/w/index.php?title=File:Flame-spray-synthesis-2.JPG License: Creative Commons Attribution-Share Alike Contributors: Evagelos Athanassiou File:Hard disk.jpg Source: http://en.wikipedia.org/w/index.php?title=File:Hard_disk.jpg License: Creative Commons Attribution-Sharealike 3.0,2.5,2.0,1.0 Contributors: Inklein File:Ferritin.png Source: http://en.wikipedia.org/w/index.php?title=File:Ferritin.png License: GNU General Public License Contributors: Simonxag, Vossman File:EasyAxis.png Source: http://en.wikipedia.org/w/index.php?title=File:EasyAxis.png License: Creative Commons Zero Contributors: RockMagnetist image:Hexagonal lattice.svg Source: http://en.wikipedia.org/w/index.php?title=File:Hexagonal_lattice.svg License: GNU Free Documentation License Contributors: Original PNGs by Daniel Mayer, traced in Inkscape by User:Stannered image:Tetragonal.png Source: http://en.wikipedia.org/w/index.php?title=File:Tetragonal.png License: GNU Free Documentation License Contributors: Daniel Mayer, DrBob image:Rhombohedral.svg Source: http://en.wikipedia.org/w/index.php?title=File:Rhombohedral.svg License: GNU Free Documentation License Contributors: Created by Daniel Mayer, traced by User:Stannered File:EasyCone.png Source: http://en.wikipedia.org/w/index.php?title=File:EasyCone.png License: Creative Commons Zero Contributors: RockMagnetist File:MagneticAnisotropyCubicAnisotropyPositive.png Source: http://en.wikipedia.org/w/index.php?title=File:MagneticAnisotropyCubicAnisotropyPositive.png License: Creative Commons Zero Contributors: RockMagnetist File:MagneticAnisotropyCubicAnisotropyNegative.png Source: http://en.wikipedia.org/w/index.php?title=File:MagneticAnisotropyCubicAnisotropyNegative.png License: Creative Commons Zero Contributors: RockMagnetist Image:Configuration of a Lithium Atom.png Source: http://en.wikipedia.org/w/index.php?title=File:Configuration_of_a_Lithium_Atom.png License: Creative Commons Attribution-Sharealike 3.0 Contributors: Aleksa Lukic Image:Klechkowski rule 2.svg Source: http://en.wikipedia.org/w/index.php?title=File:Klechkowski_rule_2.svg License: Creative Commons Attribution-ShareAlike 3.0 Unported Contributors: Sharayanan Image:Periodic Table structure.svg Source: http://en.wikipedia.org/w/index.php?title=File:Periodic_Table_structure.svg License: Creative Commons Attribution-Share Alike Contributors: Sch0013r Image:Exchangebias.png Source: http://en.wikipedia.org/w/index.php?title=File:Exchangebias.png License: Creative Commons Attribution-ShareAlike 3.0 Unported Contributors: User:Chaiken

260

License

261

License
Creative Commons Attribution-Share Alike 3.0 Unported //creativecommons.org/licenses/by-sa/3.0/

Potrebbero piacerti anche