Absolute zero

Source: Wikipedia, the free encyclopedia.

Zero kelvin (−273.15 °C) is defined as absolute zero.

Absolute zero is the lowest limit of the thermodynamic temperature scale; a state at which the enthalpy and entropy of a cooled ideal gas reach their minimum value, taken as zero kelvin. The fundamental particles of nature have minimum vibrational motion, retaining only quantum mechanical, zero-point energy-induced particle motion. The theoretical temperature is determined by extrapolating the ideal gas law; by international agreement, absolute zero is taken as −273.15 degrees on the Celsius scale (International System of Units),[1][2] which equals −459.67 degrees on the Fahrenheit scale (United States customary units or imperial units).[3] The corresponding Kelvin and Rankine temperature scales set their zero points at absolute zero by definition.

It is commonly thought of as the lowest temperature possible, but it is not the lowest enthalpy state possible, because all real substances begin to depart from the ideal gas when cooled as they approach the change of state to liquid, and then to solid; and the sum of the enthalpy of vaporization (gas to liquid) and enthalpy of fusion (liquid to solid) exceeds the ideal gas's change in enthalpy to absolute zero. In the quantum-mechanical description, matter at absolute zero is in its ground state, the point of lowest internal energy.

The

asymptotically.[4] Even a system at absolute zero, if it could somehow be achieved, would still possess quantum mechanical zero-point energy, the energy of its ground state at absolute zero; the kinetic energy
of the ground state cannot be removed.

Scientists and technologists routinely achieve temperatures close to absolute zero, where matter exhibits quantum effects such as Bose–Einstein condensate, superconductivity and superfluidity.

Thermodynamics near absolute zero

At temperatures near 0 K (−273.15 °C; −459.67 °F), nearly all molecular motion ceases and ΔS = 0 for any adiabatic process, where S is the entropy. In such a circumstance, pure substances can (ideally) form perfect crystals with no structural imperfections as T → 0. Max Planck's strong form of the third law of thermodynamics states the entropy of a perfect crystal vanishes at absolute zero. The original Nernst heat theorem makes the weaker and less controversial claim that the entropy change for any isothermal process approaches zero as T → 0:

The implication is that the entropy of a perfect crystal approaches a constant value. An adiabat is a state with constant entropy, typically represented on a graph as a curve in a manner similar to isotherms and isobars.

The

adiabat S = 0, although other isotherms and adiabats are distinct. As no two adiabats intersect, no other adiabat can intersect
the T = 0 isotherm. Consequently no adiabatic process initiated at nonzero temperature can lead to zero temperature. (≈ Callen, pp. 189–190)

A perfect crystal is one in which the internal lattice structure extends uninterrupted in all directions. The perfect order can be represented by translational symmetry along three (not usually orthogonal) axes. Every lattice element of the structure is in its proper place, whether it is a single atom or a molecular grouping. For substances that exist in two (or more) stable crystalline forms, such as diamond and graphite for carbon, there is a kind of chemical degeneracy. The question remains whether both can have zero entropy at T = 0 even though each is perfectly ordered.

Perfect crystals never occur in practice; imperfections, and even entire amorphous material inclusions, can and do get "frozen in" at low temperatures, so transitions to more stable states do not occur.

Using the Debye model, the specific heat and entropy of a pure crystal are proportional to T 3, while the enthalpy and chemical potential are proportional to T 4. (Guggenheim, p. 111) These quantities drop toward their T = 0 limiting values and approach with zero slopes. For the specific heats at least, the limiting value itself is definitely zero, as borne out by experiments to below 10 K. Even the less detailed Einstein model shows this curious drop in specific heats. In fact, all specific heats vanish at absolute zero, not just those of crystals. Likewise for the coefficient of thermal expansion. Maxwell's relations show that various other quantities also vanish. These phenomena were unanticipated.

Since the relation between changes in Gibbs free energy (G), the enthalpy (H) and the entropy is

thus, as T decreases, ΔG and ΔH approach each other (so long as ΔS is bounded). Experimentally, it is found that all spontaneous processes (including

endothermic
reactions can proceed spontaneously if the TΔS term is large enough.

Moreover, the slopes of the derivatives of ΔG and ΔH converge and are equal to zero at T = 0. This ensures that ΔG and ΔH are nearly the same over a considerable range of temperatures and justifies the approximate empirical Principle of Thomsen and Berthelot, which states that the equilibrium state to which a system proceeds is the one that evolves the greatest amount of heat, i.e., an actual process is the most exothermic one. (Callen, pp. 186–187)

One model that estimates the properties of an

velocities, even at absolute zero. The maximum energy that electrons can have at absolute zero is called the Fermi energy. The Fermi temperature is defined as this maximum energy divided by the Boltzmann constant, and is on the order of 80,000 K for typical electron densities found in metals. For temperatures significantly below the Fermi temperature, the electrons behave in almost the same way as at absolute zero. This explains the failure of the classical equipartition theorem
for metals that eluded classical physicists in the late 19th century.

Relation with Bose–Einstein condensate

Velocity-distribution data of a gas of rubidium atoms at a temperature within a few billionths of a degree above absolute zero. Left: just before the appearance of a Bose–Einstein condensate. Center: just after the appearance of the condensate. Right: after further evaporation, leaving a sample of nearly pure condensate.

A Bose–Einstein condensate (BEC) is a state of matter of a dilute gas of weakly interacting bosons confined in an external potential and cooled to temperatures very near absolute zero. Under such conditions, a large fraction of the bosons occupy the lowest quantum state of the external potential, at which point quantum effects become apparent on a macroscopic scale.[5]

This state of matter was first predicted by

quantum statistics of light quanta (now called photons). Einstein was impressed, translated the paper from English to German and submitted it for Bose to the Zeitschrift für Physik, which published it. Einstein then extended Bose's ideas to material particles (or matter) in two other papers.[6]

Seventy years later, in 1995, the first gaseous

nanokelvin (nK)[7] (1.7×10−7 K).[8]

A record cold temperature of 450 ± 80 picokelvin (pK) (4.5×10−10 K) in a BEC of sodium atoms was achieved in 2003 by researchers at the

black-body
(peak emittance) wavelength of 6,400 kilometers is roughly the radius of Earth.

Absolute temperature scales

Absolute, or

]

Negative temperatures

Temperatures that are expressed as negative numbers on the familiar Celsius or Fahrenheit scales are simply colder than the zero points of those scales. Certain systems can achieve truly negative temperatures; that is, their thermodynamic temperature (expressed in kelvins) can be of a negative quantity. A system with a truly negative temperature is not colder than absolute zero. Rather, a system with a negative temperature is hotter than any system with a positive temperature, in the sense that if a negative-temperature system and a positive-temperature system come in contact, heat flows from the negative to the positive-temperature system.[10]

Most familiar systems cannot achieve negative temperatures because adding energy always increases their entropy. However, some systems have a maximum amount of energy that they can hold, and as they approach that maximum energy their entropy actually begins to decrease. Because temperature is defined by the relationship between energy and entropy, such a system's temperature becomes negative, even though energy is being added.[10] As a result, the Boltzmann factor for states of systems at negative temperature increases rather than decreases with increasing state energy. Therefore, no complete system, i.e. including the electromagnetic modes, can have negative temperatures, since there is no highest energy state,[citation needed] so that the sum of the probabilities of the states would diverge for negative temperatures. However, for quasi-equilibrium systems (e.g. spins out of equilibrium with the electromagnetic field) this argument does not apply, and negative effective temperatures are attainable.

On 3 January 2013, physicists announced that for the first time they had created a quantum gas made up of potassium atoms with a negative temperature in motional degrees of freedom.[11]

History

Robert Boyle pioneered the idea of an absolute zero.

One of the first to discuss the possibility of an absolute minimal temperature was

nitre. But all of them seemed to agree that, "There is some body or other that is of its own nature supremely cold and by participation of which all other bodies obtain that quality."[13]

Limit to the "degree of cold"

The question of whether there is a limit to the degree of coldness possible, and, if so, where the zero must be placed, was first addressed by the French physicist Guillaume Amontons in 1703, in connection with his improvements in the air thermometer. His instrument indicated temperatures by the height at which a certain mass of air sustained a column of mercury—the pressure, or "spring" of the air varying with temperature. Amontons therefore argued that the zero of his thermometer would be that temperature at which the spring of the air was reduced to nothing.[14] He used a scale that marked the boiling point of water at +73 and the melting point of ice at +51+12, so that the zero was equivalent to about −240 on the Celsius scale.[15] Amontons held that the absolute zero cannot be reached, so never attempted to compute it explicitly.[16] The value of −240 °C, or "431 divisions [in Fahrenheit's thermometer] below the cold of freezing water"[17] was published by George Martine in 1740.

This close approximation to the modern value of −273.15 °C[1] for the zero of the air thermometer was further improved upon in 1779 by Johann Heinrich Lambert, who observed that −270 °C (−454.00 °F; 3.15 K) might be regarded as absolute cold.[18]

Values of this order for the absolute zero were not, however, universally accepted about this period. Pierre-Simon Laplace and Antoine Lavoisier, in their 1780 treatise on heat, arrived at values ranging from 1,500 to 3,000 below the freezing point of water, and thought that in any case it must be at least 600 below. John Dalton in his Chemical Philosophy gave ten calculations of this value, and finally adopted −3,000 °C as the natural zero of temperature.

Charles's law

From 1787 to 1802, it was determined by Jacques Charles (unpublished), John Dalton,[19] and Joseph Louis Gay-Lussac[20] that, at constant pressure, ideal gases expanded or contracted their volume linearly (Charles's law) by about 1/273 parts per degree Celsius of temperature's change up or down, between 0° and 100° C. This suggested that the volume of a gas cooled at about −273 °C would reach zero.

Lord Kelvin's work

After

Lord Kelvin approached the question from an entirely different point of view, and in 1848 devised a scale of absolute temperature that was independent of the properties of any particular substance and was based on Carnot's theory of the Motive Power of Heat and data published by Henri Victor Regnault.[21] It followed from the principles on which this scale was constructed that its zero was placed at −273 °C, at almost precisely the same point as the zero of the air thermometer,[15] where the air volume would reach "nothing". This value was not immediately accepted; values ranging from −271.1 °C (−455.98 °F) to −274.5 °C (−462.10 °F), derived from laboratory measurements and observations of astronomical refraction, remained in use in the early 20th century.[22]

The race to absolute zero

Commemorative plaque in Leiden

With a better theoretical understanding of absolute zero, scientists were eager to reach this temperature in the lab.

Zygmunt Wróblewski and Karol Olszewski
.

Scottish chemist and physicist

superfluids
for the first time.

Very low temperatures

The rapid expansion of gases leaving the Boomerang Nebula, a bi-polar, filamentary, likely proto-planetary nebula in Centaurus, has a temperature of 1 K, the lowest observed outside of a laboratory.

The average temperature of the universe today is approximately 2.73 kelvins (−454.76 °F), or about −270.42 °C, based on measurements of

electron temperature (total energy divided by particle count) which has increased over time.[29]

Absolute zero cannot be achieved, although it is possible to reach temperatures close to it through the use of

phenomena
, scientists have worked to obtain even lower temperatures.

See also

References

  1. ^ a b c "SI Brochure: The International System of Units (SI) – 9th edition (updated in 2022)". BIPM. p. 133. Retrieved 7 September 2022. [...], it remains common practice to express a thermodynamic temperature, symbol T, in terms of its difference from the reference temperature T0 = 273.15 K, close to the ice point. This difference is called the Celsius temperature
  2. .
  3. ^ Zielinski, Sarah (1 January 2008). "Absolute Zero". Smithsonian Institution. Archived from the original on 1 April 2013. Retrieved 26 January 2012.
  4. PMID 28290452
  5. .
  6. ^ "New State of Matter Seen Near Absolute Zero". NIST. Archived from the original on 1 June 2010.
  7. ^ Levi, Barbara Goss (2001). "Cornell, Ketterle, and Wieman Share Nobel Prize for Bose–Einstein Condensates". Search & Discovery. Physics Today online. Archived from the original on 24 October 2007. Retrieved 26 January 2008.
  8. (PDF) from the original on 9 October 2022.
  9. ^ a b Chase, Scott. "Below Absolute Zero -What Does Negative Temperature Mean?". The Physics and Relativity FAQ. Archived from the original on 15 August 2011. Retrieved 2 July 2010.
  10. S2CID 124101032
    .
  11. ^ Stanford, John Frederick (1892). The Stanford Dictionary of Anglicised Words and Phrases.
  12. ^ Boyle, Robert (1665). New Experiments and Observations touching Cold.
  13. ^ Amontons (18 April 1703). "Le thermomètre rèduit à une mesure fixe & certaine, & le moyen d'y rapporter les observations faites avec les anciens Thermométres" [The thermometer reduced to a fixed & certain measurement, & the means of relating to it observations made with old thermometers]. Histoire de l'Académie Royale des Sciences, avec les Mémoires de Mathématique et de Physique pour la même Année (in French): 50–56. Amontons described the relation between his new thermometer (which was based on the expansion and contraction of alcohol (esprit de vin)) and the old thermometer (which was based on air). From p. 52: " […] d'où il paroît que l'extrême froid de ce Thermométre seroit celui qui réduiroit l'air à ne soutenir aucune charge par son ressort, […] " ([…] whence it appears that the extreme cold of this [air] thermometer would be that which would reduce the air to supporting no load by its spring, […]) In other words, the lowest temperature which can be measured by a thermometer which is based on the expansion and contraction of air is that temperature at which the air's pressure ("spring") has decreased to zero.
  14. ^ a b Chisholm, Hugh, ed. (1911). "Cold" . Encyclopædia Britannica (11th ed.). Cambridge University Press.
  15. .
  16. ^ Martine, George (1740). "Essay VI: The various degrees of heat in bodies". Essays Medical and Philosophical. London, England, UK: A. Millar. p. 291.
  17. OCLC 165756016
    .
  18. ^ J. Dalton (1802), "Essay II. On the force of steam or vapour from water and various other liquids, both in vacuum and in air" and Essay IV. "On the expansion of elastic fluids by heat," Memoirs of the Literary and Philosophical Society of Manchester, vol. 8, pt. 2, pp. 550–74, 595–602.
  19. ^ Gay-Lussac, J. L. (1802), "Recherches sur la dilatation des gaz et des vapeurs", Annales de Chimie, XLIII: 137. English translation (extract).
  20. ^ Thomson, William (1848). "On an Absolute Thermometric Scale founded on Carnot's Theory of the Motive Power of Heat, and calculated from Regnault's observations". Proceedings of the Cambridge Philosophical Society. 1: 66–71.
  21. OCLC 64423127
  22. ^ "ABSOLUTE ZERO – PBS NOVA DOCUMENTARY (full length)". YouTube. Archived from the original on 6 April 2017. Retrieved 23 November 2016.
  23. ^ Cryogenics. Scienceclarified.com. Retrieved on 22 July 2012.
  24. ^ "The Nobel Prize in Physics 1913: Heike Kamerlingh Onnes". Nobel Media AB. Retrieved 24 April 2012.
  25. ^ Kruszelnicki, Karl S. (25 September 2003). "Coldest Place in the Universe 1". Australian Broadcasting Corporation. Retrieved 24 September 2012.
  26. ^ "What's the temperature of space?". The Straight Dope. 3 August 2004. Retrieved 24 September 2012.
  27. S2CID 238730757
    .
  28. ^ "History of temperature changes in the Universe revealed—First measurement using the Sunyaev-Zeldovich effect". Kavli Institute for the Physics and Mathematics of the Universe. 10 November 2020.
  29. S2CID 244005391
    .
  30. ^ Catchpole, Heather (4 September 2008). "Cosmos Online – Verging on absolute zero". Archived from the original on 22 November 2008.
  31. ISBN 978-951-22-5208-4. Archived from the original
    on 28 April 2001. Retrieved 11 February 2008.
  32. ^ "Low Temperature World Record" (Press release). Low Temperature Laboratory, Teknillinen Korkeakoulu. 8 December 2000. Archived from the original on 18 February 2008. Retrieved 11 February 2008.
  33. S2CID 121465475
    .
  34. ^ "Mysterious Sedna | Science Mission Directorate". science.nasa.gov. Retrieved 25 November 2022.
  35. ^ "Scientific Perspectives for ESA's Future Programme in Life and Physical sciences in Space" (PDF). esf.org. Archived from the original (PDF) on 6 October 2014. Retrieved 28 March 2014.
  36. ^ "Atomic Quantum Sensors in Space" (PDF). University of California, Los Angeles. Archived (PDF) from the original on 9 October 2022.
  37. ^ "Atoms Reach Record Temperature, Colder than Absolute Zero". livescience.com. 3 January 2013.
  38. ^ "CUORE: The Coldest Heart in the Known Universe". INFN Press Release. Retrieved 21 October 2014.
  39. ^ "MIT team creates ultracold molecules". Massachusetts Institute of Technology, Massachusetts, Cambridge. Archived from the original on 18 August 2015. Retrieved 10 June 2015.
  40. ^ "Coolest science ever headed to the space station". Science | AAAS. 5 September 2017. Retrieved 24 September 2017.
  41. ^ "Cold Atom Laboratory Mission". Jet Propulsion Laboratory. NASA. 2017. Archived from the original on 29 March 2013. Retrieved 22 December 2016.
  42. ^ "Cold Atom Laboratory Creates Atomic Dance". NASA News. 26 September 2014. Retrieved 21 May 2015.
  43. S2CID 237396804
    .

Further reading

External links