Jump to content

Electromotive force

From Wikipedia, the free encyclopedia

This is an old revision of this page, as edited by 24.150.150.55 (talk) at 00:14, 21 December 2010 (→‎Notation and units of measurement). The present address (URL) is a permanent link to this revision, which may differ significantly from the current revision.

This article is about the physics term. For the band, see EMF (band).
for the provider of electric vehicle recharging stations, see Elektromotive.

In physics, electromotive force, or most commonly emf (seldom capitalized), or (occasionally) electromotance is "that which tends to cause current (actual electrons and ions) to flow."[1]

More formally, emf is the external work expended per unit of charge to produce an electric potential difference across two open-circuited terminals.[2][3] The electric potential difference is created by separating positive and negative charges, thereby generating an electric field.[4][5] The created electrical potential difference drives current flow if a circuit is attached to the source of emf. When current flows, however, the voltage across the terminals of the source of emf is no longer the open-circuit value, due to voltage drops inside the device due to its internal resistance.

Devices that can provide emf include voltaic cells, thermoelectric devices, solar cells, electrical generators, transformers, and even Van de Graaff generators.[2][6] In nature, emf is generated whenever magnetic field fluctuations occur through a surface. An example for this is the varying Earth magnetic field during a geomagnetic storm, acting on anything on the surface of the planet, like an extended electrical grid.

In the case of a battery, charge separation that gives rise to a voltage difference is accomplished by chemical reactions at the electrodes;[5] a voltaic cell can be thought of as having a "charge pump" of atomic dimensions at each electrode, that is:

"A source of emf can be thought of as a kind of charge pump that acts to move positive charge from a point of low potential through its interior to a point of high potential. … By chemical, mechanical or other means, the source of emf performs work dW on that charge to move it to the high potential terminal. The emf of the source is defined as the work dW done per charge dq: = dW/dq." [7]

The reactions at the electrode–electrolyte interfaces provide the "seat" of emf for the voltaic cell, that is, these reactions drive the current.[8] In the open-circuit case, charge separation continues until the electrical field from the separated charges is sufficient to arrest the reactions.

In the case of an electrical generator, a time-varying magnetic field inside the generator creates an electric field via electromagnetic induction, which in turn creates an energy difference between generator terminals. Charge separation takes place within the generator, with electrons flowing away from one terminal and toward the other, until, in the open-circuit case, sufficient electric field builds up to make further movement unfavorable. Again the emf is countered by the electrical voltage due to charge separation. If a load is attached, this voltage can drive a current. The general principle governing the emf in such electrical machines is Faraday's law of induction.

A solar cell or photodiode is another source of emf, with light energy as the external power source.

Notation and units of measurement

Electromotive force is often denoted by or (script capital E, Unicode U+2130).

In a device without internal resistance, if an electric charge Q passes through that device, and gains an energy W, the net emf for that device is the energy gained per unit charge, or W/Q. Like other measures of energy per charge, emf has SI units of volts, equivalent to joules per coulomb.[9]

Electromotive force in electrostatic units is the statvolt (in the centimeter gram second system of units equal in amount to an erg per electrostatic unit of charge).

Terminology

The term electromotive force is due to Alessandro Volta (1745–1827), who invented the battery, or voltaic pile. "Electromotive force" originally referred to the 'force' with which positive and negative charges could be separated (that is, moved, hence "electromotive"), and was also called "electromotive power" (although it is not a power in the modern sense). Maxwell's 1865 explanation of what are now called Maxwell's equations used the term "electromotive force" for what is now called the electric field strength.[10] But, in his later textbook[11] he uses the term "electromotive force" both for "voltage-like" causes of current flow in an electric circuit, and (inconsistently) for contact potential difference (which is a form of electrostatic potential difference). Given that Maxwell's textbook was written before the discovery of the electron, it is understandable that Maxwell exhibits what (in terms of modern knowledge) is inconsistency in the use of the term "electromotive force".

The word "force" in "electromotive force" is a misnomer:[12]

"[Electromotive force] has turned out to be an unfortunate choice of words which is still with us 160 years later. In all of physics except electromagnetic induction, the term 'force' is reserved for mechanical action on ponderable matter and is measured in units called Newtons. In contrast electromotive force is measured in units of Volts and causes charge separation."[12]

Nonetheless, the term "electromotive force" has resisted change. "Electromotance", meaning (literally) tendency to move ("-motance") electrical charge, is semantically more accurate, but not widely adopted. Both terms are less common than the abbreviation emf.

These terms (emf, voltage, etc.) have many interpretations and applications, not all necessarily consistent with each other. The emf is typically considered to be the work done per unit charge by a source in creating a separation of positive from negative charges, thereby creating a voltage difference; the work done per unit charge in pushing charge through a battery creating the battery's voltage difference, for example. However, there is not complete unanimity upon this usage. As Sydney Ross says, in excusing himself for avoiding the term emf:[13]

We have refrained from using the term 'electromotive force' or 'e.m.f.' for short; for there is no consistency between different authors in the meaning of the term. … To some authors it is synonymous with 'voltage.' To others it means the open-circuit voltage of a battery. To a third group of authors it means the open-circuit voltage of any two-terminal device. This use is met most often in connection with Thevenin's theorem in circuit theory. To a fourth group it means the work accounted for by agencies other than differences of the (not measurable) Galvani potentials. Such authors equate the current–resistance product of a circuit branch to the sum of voltage plus e.m.f. A fifth group extends this use to field theory. The authors of this group equate the product of current density and resistivity to the sum of electric-field strength plus an e.m.f. gradient. A sixth group applies the term to electromagnetic induction. These authors define e.m.f. as the spatial line integral of the electric-field strength taken over a complete loop. To them the term 'counter e.m.f.' means something.

It is common in some fields, such as circuit theory, to refer to the voltage created by the emf as the emf.[14][15][16] Some authors do not distinguish between the emf and the voltage it creates.[17][18] Some use emf to refer to the open-circuit voltage and voltage to the potential difference when current is drawn.[19] Here is a quotation describing emf as an open-circuit voltage difference:[20]

"This buildup of charge on the electrodes tends to oppose the current flow with a 'back voltage' ΔV. On an open circuit, I = 0 the value of ΔV for which I = 0 is defined as the emf of the cell. That is, = ΔVI=0."

This usage does not identify the work done per unit charge in creating the charge build-up as emf, but rather identifies emf with the consequent "back voltage" that arrests current flow in the open-circuit condition.[21] One emphasizes the conversion of energy from other forms to electrical energy, the other emphasizes the resulting electrical potential. This article focuses upon the conversion of other forms of energy to electrical potential energy, and describes some examples of how this conversion comes about.

Formal definitions of electromotive force

Inside a source of emf that is open-circuited, the conservative electrostatic field created by separation of charge exactly cancels the forces producing the emf. Thus, the emf has the same value but opposite sign as the integral of the electric field aligned with an internal path between two terminals A and B of a source of emf in open-circuit condition (the path is taken from the negative terminal to the positive terminal to yield a positive emf, indicating work done on the electrons moving in the circuit).[22] Mathematically:

where Ecs is the conservative electrostatic field created by the charge separation associated with the emf, d is an element of the path from terminal A to terminal B, and ‘·’ denotes the vector dot product.[23] This equation applies only to locations A and B that are terminals, and does not apply to paths between points A and B with portions outside the source of emf. This equation involves the electrostatic electric field due to charge separation Ecs and does not involve (for example) any non-conservative component of electric field due to Faraday's law of induction.

In the case of a closed path in the presence of a varying magnetic field, the integral of the electric field around a closed loop may be nonzero; one common application of the concept of emf, known as "induced emf" is the voltage induced in a such a loop.[24] The "induced emf" around a stationary closed path C is:

where now E is the entire electric field, conservative and non-conservative, and the integral is around an arbitrary but stationary closed curve C through which there is a varying magnetic field. Note that the electrostatic field does not contribute to the net emf around a circuit because the electrostatic portion of the electric field is conservative (that is, the work done against the field around a closed path is zero).

This definition can be extended to arbitrary sources of emf and moving paths C:[25]

which is a conceptual equation mainly, because the determination of the "effective forces" is difficult.

Electromotive force in thermodynamics

When multiplied by an amount of charge dZ the emf ℰ yields a thermodynamic work term ℰdZ that is used in the formalism for the change in Gibbs free energy when charge is passed in a battery:

where G is the Gibb's free energy, S is the entropy, V is the system volume, P is its pressure and T is its absolute temperature.

The combination ( ℰ , Z ) is an example of a conjugate pair of variables. At constant pressure the above relationship produces a Maxwell relation that links the change in open cell voltage with temperature T (a measurable quantity) to the change in entropy S when charge is passed isothermally and isobarically. The latter is closely related to the reaction entropy of the electrochemical reaction that lends the battery its power. This Maxwell relation is:[26]

If a mole of ions goes into solution (for example, in a Daniell cell, as discussed below) the charge through the external circuit is:

where n0 is the number of electrons/ion, and F0 is the Faraday constant and the minus sign indicates discharge of the cell. Assuming constant pressure and volume, the thermodynamic properties of the cell are related strictly to the behavior of its emf by:[26]

where ΔH is the heat of reaction. The quantities on the right all are directly measurable.

Electromotive force and voltage difference

An electrical voltage difference is sometimes called an emf.[14][15][16][17][19] The points below illustrate the more formal usage, in terms of the distinction between emf and the voltage it generates:

  1. For a circuit as a whole, such as one containing a resistor in series with a voltaic cell, electrical voltage does not contribute to the overall emf, because the voltage difference on going around a circuit is zero. (The ohmic IR drop plus the applied electrical voltage is zero. See Kirchhoff's Law). The emf is due solely to the chemistry in the battery that causes charge separation, which in turn creates an electrical voltage that drives the current.
  2. For a circuit consisting of an electrical generator that drives current through a resistor, the emf is due solely to a time-varying magnetic field that generates an electrical voltage that in turn drives the current. (The ohmic IR drop plus the applied electrical voltage again is zero. See Kirchhoff's Law)
  3. A transformer coupling two circuits may be considered a source of emf for one of the circuits, just as if it were caused by an electrical generator; this example illustrates the origin of the term "transformer emf".
  4. A photodiode or solar cell may be considered as a source of emf, similar to a battery, resulting in an electrical voltage generated by charge separation driven by light rather than chemical reaction.[27]
  5. Other devices that produce emf are fuel cells, thermocouples, and thermopiles.[28]

In the case of an open circuit, the electric charge that has been separated by the mechanism generating the emf creates an electric field opposing the separation mechanism. For example, the chemical reaction in a voltaic cell stops when the opposing electric field at each electrode is strong enough to arrest the reactions. A larger opposing field can reverse the reactions in what are called reversible cells.[29][30]

The electric charge that has been separated creates an electric potential difference that can be measured with a voltmeter between the terminals of the device. The magnitude of the emf for the battery (or other source) is the value of this 'open circuit' voltage. When the battery is charging or discharging, the emf itself cannot be measured directly using the external voltage because some voltage is lost inside the source.[15] It can, however, be inferred from a measurement of the current I and voltage difference V, provided that the internal resistance r already has been measured: I = ( − V)/r.

Electromotive force generation

Chemical sources

A typical reaction path requires the initial reactants to cross an energy barrier, enter an intermediate state and finally emerge in a lower energy configuration. If charge separation is involved, this energy difference can result in an emf. See Bergmann et al.[31] and Transition state.
Galvanic cell using a salt bridge

The question of how batteries (galvanic cells) generate an emf is one that occupied scientists for most of the 19th century. The "seat of the electromotive force" was eventually determined by Walther Nernst to be primarily at the interfaces between the electrodes and the electrolyte.[8]

Molecules are groups of atoms held together by chemical bonds, and these bonds consist of electrical forces between electrons (negative) and protons (positive). The molecule in isolation is a stable entity, but when different molecules are brought together, some types of molecules are able to steal electrons from others, resulting in charge separation. This redistribution of charge is accompanied by a change in energy of the system, and a reconfiguration of the atoms in the molecules.[32] The gain of an electron is termed "reduction" and the loss of an electron is termed "oxidation". Reactions in which such electron exchange occurs (which are the basis for batteries) are called reduction-oxidation reactions or redox reactions. In a battery, one electrode is composed of material that gains electrons from the solute, and the other electrode loses electrons, because of these fundamental molecular attributes. The same behavior can be seen in atoms themselves, and their ability to steal electrons is referred to as their electronegativity.[33]

As an example, a Daniell cell consists of a zinc anode (an electron collector), which dissolves into a zinc sulfate solution, the dissolving zinc leaving behind its electrons in the electrode according to the oxidation reaction (s = solid electrode; aq = aqueous solution):

The zinc sulfate is an electrolyte, that is, a solution in which the components consist of ions, in this case zinc ions , and sulfate ions .

At the cathode, the copper ions in a copper sulfate electrolyte adopt electrons from the electrode by the reduction reaction:

and the thus-neutralized copper plates onto the electrode. (A detailed discussion of the microscopic process of electron transfer between an electrode and the ions in an electrolyte may be found in Conway.)[34]

The electrons pass through the external circuit (light bulb in figure), while the ions pass through the salt bridge to maintain charge balance. In the process the zinc anode is dissolved while the copper electrode is plated with copper.[35] If the light bulb is removed (open circuit) the emf between the electrodes is opposed by the electric field due to charge separation, and the reactions stop.

At 273 K, the emf = 1.0934 V, with a temperature coefficient of d/dT = −4.53×10−4 V/K.[26]

Voltaic cells

Volta developed the voltaic cell about 1792, and presented his work March 20, 1800.[36] Volta correctly identified the role of dissimilar electrodes in producing the voltage, but incorrectly dismissed any role for the electrolyte.[37] Volta ordered the metals in a 'tension series', “that is to say in an order such that any one in the list becomes positive when in contact with any one that succeeds, but negative by contact with any one that precedes it.”[38] A typical symbolic convention in a schematic of this circuit ( –||– ) would have a long electrode 1 and a short electrode 2, to indicate that electrode 1 dominates. Volta's law about opposing electrode emfs means that, given ten electrodes (for example, zinc and nine other materials), which can be used to produce 45 types of voltaic cells (10 × 9/2), only nine relative measurements (for example, copper and each of the nine others) are needed to get all 45 possible emfs that these ten electrodes can produce.[citation needed]

Electromotive force of cells

The electromotive force produced by primary and secondary cells is usually of the order of a few volts. The figures quoted below are nominal, because emf varies according to the size of the load and the state of exhaustion of the cell.

Emf Cell chemistry
1.2 V nickel-cadmium
1.2 V nickel-metal hydride
1.5 V zinc-carbon
2.1 V lead-acid
3.6 V to 3.7 V lithium-ion

Electromagnetic induction

The principle of electromagnetic induction, noted above, states that a time-dependent magnetic field produces a circulating electric field. A time-dependent magnetic field can be produced either by motion of a magnet relative to a circuit, by motion of a circuit relative to another circuit (at least one of these must be carrying a current), or by changing the current in a fixed circuit. The effect on the circuit itself, of changing the current, is known as self-induction; the effect on another circuit is known as mutual induction.

For a given circuit, the electromagnetically induced emf is determined purely by the rate of change of the magnetic flux through the circuit according to Faraday's law of induction.

An emf is induced in a coil or conductor whenever there is change in the flux linkages. Depending on the way in which the changes are brought about, there are two types: When the conductor is moved in a stationary magnetic field to procure a change in the flux linkage, the emf is statically induced. The electromotive force generated by motion is often referred to as motional emf. When the change in flux linkage arises from a change in the magnetic field around the stationary conductor, the emf is dynamically induced. The electromotive force generated by a time-varying magnetic field is often referred to as transformer emf.

Contact potentials

When two different solids are in contact, it is common that thermodynamic equilibrium requires one of the solids assume a higher electrical potential than the other, the contact potential.[39] For example, dissimilar metals in contact produce what is known also as a contact electromotive force or Galvani potential. The magnitude of this potential difference often is expressed as a difference in Fermi levels in the two solids, where the Fermi level (a name for the chemical potential of an electron system[40][41]) describes the energy necessary to remove an electron from the body.[42] Evidently, if there is an energy advantage in taking an electron from one body to the other, this transfer will occur, thereby causing a charge separation, one body gaining electrons and the other losing electrons. This charge transfer causes a potential difference between the bodies, and therefore, charge transfer becomes more difficult as the charge separation increases. At thermodynamic equilibrium, the gain in energy due to Fermi level difference is matched by the work needed to surmount this potential difference, and at this point no more transfer occurs, and the potential difference has the value called the contact potential. The difference in Fermi levels, on the other hand, is referred to as the emf.[43] The contact potential cannot drive current through a load attached to its terminals because that current would involve a charge transfer. No mechanism exists to continue such transfer and, hence, maintain a current, once equilibrium is attained.

One might inquire why the contact potential does not appear in Kirchhoff's law of voltages as one contribution to the sum of potential drops. The customary answer is that any circuit involves not only a particular diode or junction, but also all the contact potentials due to wiring and so forth around the entire circuit. The sum of all the contact potentials is zero, and so they may be ignored in Kirchhoff's law.[44][45]

Solar cell

The equivalent circuit of a solar cell; parasitic resistances are ignored in the discussion of the text.
Solar cell voltage as a function of solar cell current delivered to a load for two light-induced currents IL; currents as a ratio with reverse saturation current I0. Compare with Fig. 1.4 in Nelson.[46]

Operation of a solar cell can be understood from the equivalent circuit at right. Light, if it includes photons of sufficient energy (greater than the bandgap of the material), creates mobile electron–hole pairs in a semiconductor. Charge separation occurs because of a pre-existing electric field associated with the p-n junction in thermal equilibrium (a contact potential creates the field). This charge separation between positive holes and negative electrons across a p-n junction (a diode), yields a forward voltage, the photo voltage, between the illuminated diode terminals.[47] As has been noted earlier in the terminology section, the photo voltage is sometimes referred to as the photo emf, rather than distinguishing between the effect and the cause.

The light-induced charge separation creates a reverse current through the cell's junction (that is, not in the direction that a diode normally conducts current), and the charge separation causes a photo voltage that drives current through any attached load. However, a side effect of this voltage is that it tends to forward bias the junction. At high enough levels, this forward bias of the junction will cause a forward current in the diode that subtracts from the current created by the light. Consequently, the greatest current is obtained under short-circuit conditions, and is denoted as IL (for light-induced current) in the equivalent circuit.[48] Approximately this same current is obtained for forward voltages up to the point where the diode conduction becomes significant.

With this notation, the current-voltage relation for the illuminated diode is:

where I is the current delivered to the load, I0 is the reverse saturation current, and m the ideality factor, two parameters that depend on the solar cell construction and to some degree upon the voltage itself,[48] and where kT/q is the thermal voltage (about 0.026 V at room temperature). This relation is plotted in the figure using a fixed value m = 2.[49] Under open-circuit conditions (that is, as I → 0), the open-circuit voltage is the voltage at which forward bias of the junction is enough that the forward current completely balances the photocurrent. Rearrangement of the I–V equation provides the open-circuit voltage as:

which is useful in indicating a logarithmic dependence of Voc upon the light-induced current. Typically, the open-circuit voltage is not more than about 0.5 V.[50]

The value of the photo voltage when driving a load is variable. As shown in the figure, for a load resistance RL, the cell develops a voltage between the short-circuit value V = 0, I = IL and the open-circuit value Voc, I = 0, a value given by Ohm's law V = I RL, where the current I is the difference between the short-circuit current and current due to forward bias of the junction, as indicated by the equivalent circuit (neglecting the parasitic resistances).[46]

In contrast to the battery, at current levels near IL, the solar cell acts more like a current source rather than a voltage source.[46] The current drawn is nearly fixed over a range of load voltages, at one electron per converted photon. The quantum efficiency, or probability of getting an electron of photocurrent per incident photon, depends not only upon the solar cell itself, but upon the spectrum of the light.

The diode possesses a "built-in potential" due to the contact potential difference between the two different materials on either side of the junction. This built-in potential is established when the junction is formed as a by-product of thermodynamic equilibrium. Once established, this potential difference cannot drive a current, however, as connecting a load does not upset this equilibrium. In contrast, the accumulation of excess electrons in one region and of excess holes in another due to illumination results in a photo voltage that does drive a current when a load is attached to the illuminated diode. As noted above, this photo voltage also forward biases the junction, and so reduces the pre-existing field in the depletion region.

See also

References

  1. ^ Irving Langmuir (1916). "The Relation Between Contact Potentials and Electrochemical Action". Transactions of the American Electrochemical Society. 29. The Society: 125–182.
  2. ^ a b Lawrence M Lerner (1997). Physics for scientists and engineers. Jones & Bartlett Publishers. pp. 724–727. ISBN 0763704601.
  3. ^ David M. Cook (2003). The Theory of the Electromagnetic Field. Courier Dover. p. 157. ISBN 9780486425672.
  4. ^ Robert L. Lehrman (1998). Physics the easy way. Barron's Educational Series. p. 274. ISBN 9780764102363.
  5. ^ a b Alvin M. Halpern, Erich Erlbach (1998). Schaum's outline of theory and problems of beginning physics II. McGraw-Hill Professional. p. 138. ISBN 0070257078.
  6. ^ Paul A. Tipler and Gene Mosca (2007). Physics for Scientists and Engineers (6 ed.). Macmillan. p. 850. ISBN 142920124X.
  7. ^ Kongbam Chandramani Singh (2009). "§3.16 EMF of a source". Basic Physics. Prentice Hall India Pvt Ltd. p. 152. ISBN 8120337085.
  8. ^ a b Florian Cajori (1899). A History of Physics in Its Elementary Branches: Including the Evolution of Physical Laboratories. The Macmillan Company. pp. 218–219.
  9. ^ Van Valkenburgh (1995). Basic Electricity. Cengage Learning. pp. 1–46. ISBN 9780790610412.
  10. ^ Edward J. Rothwell and Michael J. Cloud (2001). Electromagnetics. CRC Press. p. 22. ISBN 0-8493-1397-X.
  11. ^ J. C. Maxwell (1891, republished 1998). An elementary treatise on electricity (3rd ed.). Oxford: Clarendon. {{cite book}}: Check date values in: |year= (help)CS1 maint: year (link)
  12. ^ a b Neal Graneau (2006). In the grip of the distant universe. World Scientific. p. 191. ISBN 9812567542.
  13. ^ Sydney Ross (1991). "Supplementary Note to The story of the Volta potential". Nineteenth-century attitudes: men of science. Springer. p. 83. ISBN 9780792313083.
  14. ^ a b M. Fogiel (2002). Basic Electricity. Research & Education Association. p. 76. ISBN 087891420X.
  15. ^ a b c David Halliday, Robert Resnick, and Jearl Walker (2008). Fundamentals of Physics (6th ed.). Wiley. p. 638. ISBN 9780471758013.{{cite book}}: CS1 maint: multiple names: authors list (link)
  16. ^ a b Roger L Freeman (2005). Fundamentals of Telecommunications (2nd ed.). Wiley. p. 576. ISBN 0471710458.
  17. ^ a b Terrell Croft (1917). Practical Electricity. McGraw-Hill. p. 533.
  18. ^ Vladimir Borisovich Rojansky (1979). Electromagnetic fields and waves. Courier Dover. p. 186. ISBN 9780486638348.
  19. ^ a b Leonard B Loeb (2007). Fundamentals of Electricity and Magnetism (Reprint of Wiley 1947 3rd ed.). Read Books. p. 86. ISBN 1406707333.
  20. ^ Wayne M. Saslow (2002). "§7.8 Emf and Ohm's Law". Electricity, Magnetism, and Light (3rd ed.). Academic Press. pp. 304 ff. ISBN 0126194556.
  21. ^ Note the minus sign in Eq. (7.10) in David J Griffiths (1999). Introduction to Electrodynamics (3rd ed.). Pearson/Adisson Wesley. p. 293. ISBN 013805326X.. Also, the term "back voltage" indicates the electric field due to charge separation opposes the mechanism creating the separation. In the open-circuit condition, this electric field arrests the charge separation, and zero current flows.
  22. ^ David J Griffiths (1999). Introduction to Electrodynamics (3rd ed.). Pearson/Adisson Wesley. p. 293. ISBN 013805326X.
  23. ^ Only the electric field due to the charge separation caused by the emf is counted. In a solar cell, for example, an electric field is present related to the contact potential that results from thermodynamic equilibrium (discussed later), and this electric field component is not included in the integral. Rather, only the electric field due to the particular portion of charge separation that causes the photo voltage is included.
  24. ^ Richard P. Olenick, Tom M. Apostol and David L. Goodstein (1986). Beyond the mechanical universe: from electricity to modern physics. Cambridge University Press. p. 245. ISBN 9780521304306.
  25. ^ David M. Cook (2003). The Theory of the Electromagnetic Field. Courier Dover. p. 158. ISBN 9780486425672.
  26. ^ a b c Colin B P Finn (1992). Thermal Physics. CRC Press. p. 163. ISBN 0748743790.
  27. ^ Jenny Nelson (2003). The Physics of Solar Cells. Imperial College Press. p. 6. ISBN 1860943497.
  28. ^ John S. Rigden, (editor in chief), Macmillan encyclopedia of physics. New York : Macmillan, 1996.
  29. ^ J. R. W. Warn, A. P. H. Peters (1996). Concise Chemical Thermodynamics (2 ed.). CRC Press. p. 123. ISBN 0748744452.
  30. ^ Samuel Glasstone (2007). Thermodynamics for Chemists (Reprint of D. Van Nostrand Co (1964) ed.). Read Books. p. 301. ISBN 1406773220.
  31. ^ Nikolaus Risch (2002). "Molecules - bonds and reactions". In L Bergmann; et al. (eds.). Constituents of Matter: Atoms, Molecules, Nuclei, and Particles. CRC Press. ISBN 0849312027. {{cite book}}: Explicit use of et al. in: |editor= (help)
  32. ^ The brave reader can find an extensive discussion for organic electrochemistry in Christian Amatore (2000). "Basic concepts". In Henning Lund, Ole Hammerich (ed.). Organic electrochemistry (4 ed.). CRC Press. ISBN 0824704304.
  33. ^ The idea of electronegativity has been extended to include the concept of electronegativity equalization, the notion that when molecules are brought together the electrons rearrange to achieve an equilibrium where there is no net force upon them. See, for example, Francis A. Carey, Richard J. Sundberg (2007). Advanced organic chemistry (5 ed.). Springer. p. 11. ISBN 0387683461.
  34. ^ BE Conway (1999). "Energy factors in relation to electrode potential". Electrochemical supercapacitors. Springer. p. 37. ISBN 0306457369.
  35. ^ R. J. D. Tilley (2004). Understanding Solids. Wiley. p. 267. ISBN 0470852755.
  36. ^ Paul Fleury Mottelay (2008). Bibliographical History of Electricity and Magnetism (Reprint of 1892 ed.). Read Books. p. 247. ISBN 1443728446.
  37. ^ Helge Kragh (2000). "Confusion and Controversy: Nineteenth-century theories of the voltaic pile" (PDF). Nuova Voltiana:Studies on Volta and his times. Università degli studi di Pavia.
  38. ^ Linnaus Cumming (2008). An Introduction to the Theory of Electricity (Reprint of 1885 ed.). BiblioBazaar. p. 118. ISBN 0559207425.
  39. ^ George L. Trigg (1995). Landmark experiments in twentieth century physics (Reprint of Crane, Russak & Co 1975 ed.). Courier Dover. pp. 138 ff. ISBN 048628526X.
  40. ^ Angus Rockett (2007). "Diffusion and drift of carriers". Materials science of semiconductors. New York, NY: Springer Science. pp. 74 ff. ISBN 0387256539. {{cite book}}: Unknown parameter |unused_data= ignored (help)
  41. ^ Charles Kittel (2004). "Chemical potential in external fields". Elementary Statistical Physics (Reprint of Wiley 1958 ed.). Courier Dover. p. 67. ISBN 0486435148.
  42. ^ George W. Hanson (2007). Fundamentals of Nanoelectronics. Prentice Hall. p. 100. ISBN 0131957082.
  43. ^ Norio Sato (1998). "Semiconductor photoelectrodes". Electrochemistry at metal and semiconductor electrodes (2nd ed.). Elsevier. pp. 110 ff. ISBN 0444828060.
  44. ^ Richard S. Quimby (2006). Photonics and lasers. Wiley. p. 176. ISBN 0471719749.
  45. ^ Donald A. Neamen (2002). Semiconductor physics and devices (3rd ed.). McGraw-Hill Professional. p. 240. ISBN 0072321075.
  46. ^ a b c Jenny Nelson (2003). Solar cells. Imperial College Press. p. 8. ISBN 1860943497.
  47. ^ S M Dhir (2000). "§3.1 Solar cells". Electronic Components and Materials: Principles, Manufacture and Maintenance. Tata McGraw-Hill. ISBN 0074630822.
  48. ^ a b Gerardo L. Araújo (1994). "§2.5.1 Short-circuit current and open-circuit voltage". In Eduardo Lorenzo (ed.). Solar Electricity: Engineering of photovoltaic systems. Progenza for Universidad Politechnica Madrid. p. 74. ISBN 8486505550.
  49. ^ In practice, at low voltages m → 2, whereas at high voltages m → 1. See Araújo, op. cit. isbn = 8486505550. page 72
  50. ^ Robert B. Northrop (2005). "§6.3.2 Photovoltaic Cells". Introduction to Instrumentation and Measurements. CRC Press. p. 176. ISBN 0849378982.

Further reading

  • Andrew Gray, "Absolute Measurements in Electricity and Magnetism", Electromotive force. Macmillan and co., 1884.
  • John O'M. Bockris, Amulya K. N. Reddy (1973). "Electrodics". Modern Electrochemistry: An Introduction to an Interdisciplinary Area (2 ed.). Springer. ISBN 0306250020.
  • Roberts, Dana (1983). "How batteries work: A gravitational analog". Am. J. Phys. 51: 829. doi:10.1119/1.13128.
  • Charles Albert Perkins, "Outlines of Electricity and Magnetism", Measurement of Electromotive Force. Henry Holt and co., 1896.
  • John Livingston Rutgers Morgan, "The Elements of Physical Chemistry", Electromotive force. J. Wiley, 1899.
  • George F. Barker, "On the measurement of electromotive force". Proceedings of the American Philosophical Society Held at Philadelphia for Promoting Useful Knowledge, American Philosophical Society. January 19, 1883.
  • "Abhandlungen zur Thermodynamik, von H. Helmholtz. Hrsg. von Max Planck". (Tr. "Papers to thermodynamics, on H. Helmholtz. Hrsg. by Max Planck".) Leipzig, W. Engelmann, Of Ostwald classical author of the accurate sciences series. New consequence. No. 124, 1902.
  • Nabendu S. Choudhury, "Electromotive force measurements on cells involving [beta]-alumina solid electrolyte". NASA technical note, D-7322.
  • Henry S. Carhart, "Thermo-electromotive force in electric cells, the thermo-electromotive force between a metal and a solution of one of its salts". New York, D. Van Nostrand company, 1920. LCCN 20020413
  • Hazel Rossotti, "Chemical applications of potentiometry". London, Princeton, N.J., Van Nostrand, 1969. ISBN 0-442-07048-9 LCCN 69011985 //r88
  • Theodore William Richards and Gustavus Edward Behr, jr., "The electromotive force of iron under varying conditions, and the effect of occluded hydrogen". Carnegie Institution of Washington publication series , 1906. LCCN 07003935 //r88
  • G. W. Burns, et al., "Temperature-electromotive force reference functions and tables for the letter-designated thermocouple types based on the ITS-90". Gaithersburg, MD : U.S. Dept. of Commerce, National Institute of Standards and Technology, Washington, Supt. of Docs., U.S. G.P.O., 1993.
  • Norio Sato (1998). "Semiconductor photoelectrodes". Electrochemistry at metal and semiconductor electrodes (2nd ed.). Elsevier. pp. 326 ff. ISBN 0444828060.