Phase Transition
Phase Transition
58 languages
Article
Talk
Read
Edit
View history
Tools
From Wikipedia, the free encyclopedia
This diagram shows the nomenclature for the different phase transitions.
In chemistry, thermodynamics, and other related fields like physics and biology, a
phase transition (or phase change) is the physical process of transition between
one state of a medium and another. Commonly the term is used to refer to
changes among the basic states of matter: solid, liquid, and gas, and in rare
cases, plasma. A phase of a thermodynamic system and the states of matter have
uniform physical properties. During a phase transition of a given medium, certain
properties of the medium change as a result of the change of external conditions,
such as temperature or pressure. This can be a discontinuous change; for
example, a liquid may become gas upon heating to its boiling point, resulting in
an abrupt change in volume. The identification of the external conditions at which
a transformation occurs defines the phase transition point.
States of matter[edit]
A simplified phase diagram for water, showing whether solid ice, liquid water, or gaseous water
vapor is the most stable at different combinations of temperature and pressure.
Common transitions between the solid, liquid, and gaseous phases of a single
component, due to the effects of temperature and/or pressure are identified in the
following table:
From
Solid Melting Sublimation
Plasma Recombinatio
n
For a single component, the most stable phase at different temperatures and
pressures can be shown on a phase diagram. Such a diagram usually depicts
states in equilibrium. A phase transition usually occurs when the pressure or
temperature changes and the system crosses from one region to another, like
water turning from liquid to solid as soon as the temperature drops below the
freezing point. In exception to the usual case, it is sometimes possible to change
the state of a system diabatically (as opposed to adiabatically) in such a way that
it can be brought past a phase transition point without undergoing a phase
transition. The resulting state is metastable, i.e., less stable than the phase to
which the transition would have occurred, but not unstable either. This occurs in
superheating and supercooling, for example. Metastable states do not appear on
usual phase diagrams.
Structural[edit]
Phase transitions can also occur when a solid changes to a different structure
without changing its chemical makeup. In elements, this is known as allotropy,
whereas in compounds it is known as polymorphism. The change from one
crystal structure to another, from a crystalline solid to an amorphous solid, or
from one amorphous structure to another (polyamorphs) are all examples of solid
to solid phase transitions.
Magnetic[edit]
Phase transitions can also describe the change between different kinds of
magnetic ordering. The most well-known is the transition between the
ferromagnetic and paramagnetic phases of magnetic materials, which occurs at
what is called the Curie point. Another example is the transition between
differently ordered, commensurate or incommensurate, magnetic structures,
such as in cerium antimonide. A simplified but highly useful model of magnetic
phase transitions is provided by the Ising Model
Mixtures[edit]
A binary phase diagram showing the most stable chemical compounds of titanium and nickel at
different mixing ratios and temperatures.
Phase transitions involving solutions and mixtures are more complicated than
transitions involving a single compound. While chemically pure compounds
exhibit a single temperature melting point between solid and liquid phases,
mixtures can either have a single melting point, known as congruent melting, or
they have different liquidus and solidus temperatures resulting in a temperature
span where solid and liquid coexist in equilibrium. This is often the case in solid
solutions, where the two components are isostructural.
There are also a number of phase transitions involving three phases: a eutectic
transformation, in which a two-component single-phase liquid is cooled and
transforms into two solid phases. The same process, but beginning with a solid
instead of a liquid is called a eutectoid transformation. A peritectic
transformation, in which a two-component single-phase solid is heated and
transforms into a solid phase and a liquid phase. A peritectoid reaction is a
peritectoid rection, except involving only solid phases. A monotectic reaction
consists of change from a liquid and to a combination of a solid and a second
liquid, where the two liquids display a miscibility gap.[1]
Separation into multiple phases can occur via spinodal decomposition, in which a
single phase is cooled and separates into two different compositions.
Other examples[edit]
A small piece of rapidly melting solid argon shows two concurrent phase changes. The transition
from solid to liquid, and gas to liquid (shown by the white condensed water vapour).
Phase transitions occur when the thermodynamic free energy of a system is non-
analytic for some choice of thermodynamic variables (cf. phases). This condition
generally stems from the interactions of a large number of particles in a system,
and does not appear in systems that are small. Phase transitions can occur for
non-thermodynamic systems, where temperature is not a parameter. Examples
include: quantum phase transitions, dynamic phase transitions, and topological
(structural) phase transitions. In these types of systems other parameters take
the place of temperature. For instance, connection probability replaces
temperature for percolating networks.
● Phases
● Phase transition
● QCP
hide
States of matter
● Solid
● Liquid
● Gas
● Plasma
● Bose–Einstein condensate
● Bose gas
● Fermionic condensate
● Fermi gas
● Fermi liquid
● Supersolid
● Superfluidity
● Luttinger liquid
● Time crystal
show
Phase phenomena
show
Electronic phases
show
Electronic phenomena
show
Magnetic phases
show
Quasiparticles
show
Soft matter
show
Scientists
● Physics portal
● Category
● V
● T
● E
Classifications[edit]
Ehrenfest classification[edit]
In practice, only the first- and second-order phase transitions are typically
observed. The second-order phase transition was for a while controversial, as it
seems to require two sheets of the Gibbs free energy to osculate exactly, which is
so unlikely as to never occur in practice. Cornelis Gorter replied the criticism by
pointing out that the Gibbs free energy surface might have two sheets on one
side, but only one sheet on the other side, creating a forked appearance. [11] ([9] pp.
146--150)
The first example of a phase transition which did not fit into the Ehrenfest
classification was the exact solution of the Ising model, discovered in 1944 by
Lars Onsager. The exact specific heat differed from the earlier mean-field
approximations, which had predicted that it has a simple discontinuity at critical
temperature. Instead, the exact specific heat had a logarithmic divergence at the
critical temperature.[12] In the following decades, the Ehrenfest classification was
replaced by a simplified classification scheme that is able to incorporate such
transitions.
Modern classifications[edit]
In the modern classification scheme, phase transitions are divided into two broad
categories, named similarly to the Ehrenfest classes:[5]
First-order phase transitions are those that involve a latent heat. During such a
transition, a system either absorbs or releases a fixed (and typically large)
amount of energy per volume. During this process, the temperature of the system
will stay constant as heat is added: the system is in a "mixed-phase regime" in
which some parts of the system have completed the transition and others have
not.[13][14]
Familiar examples are the melting of ice or the boiling of water (the water does
not instantly turn into vapor, but forms a turbulent mixture of liquid water and
vapor bubbles). Yoseph Imry and Michael Wortis showed that quenched disorder
can broaden a first-order transition. That is, the transformation is completed over
a finite range of temperatures, but phenomena like supercooling and
superheating survive and hysteresis is observed on thermal cycling. [15][16][17]
Apart from isolated, simple phase transitions, there exist transition lines as well
as multicritical points, when varying external parameters like the magnetic field
or composition.
The liquid–glass transition is observed in many polymers and other liquids that
can be supercooled far below the melting point of the crystalline phase. This is
atypical in several respects. It is not a transition between thermodynamic ground
states: it is widely believed that the true ground state is always crystalline. Glass
is a quenched disorder state, and its entropy, density, and so on, depend on the
thermal history. Therefore, the glass transition is primarily a dynamic
phenomenon: on cooling a liquid, internal degrees of freedom successively fall
out of equilibrium. Some theoretical methods predict an underlying phase
transition in the hypothetical limit of infinitely long relaxation times. [19][20] No
direct experimental evidence supports the existence of these transitions.
Characteristic properties[edit]
Phase coexistence[edit]
A disorder-broadened first-order transition occurs over a finite range of
temperatures where the fraction of the low-temperature equilibrium phase grows
from zero to one (100%) as the temperature is lowered. This continuous variation
of the coexisting fractions with temperature raised interesting possibilities. On
cooling, some liquids vitrify into a glass rather than transform to the equilibrium
crystal phase. This happens if the cooling rate is faster than a critical cooling
rate, and is attributed to the molecular motions becoming so slow that the
molecules cannot rearrange into the crystal positions. [21] This slowing down
happens below a glass-formation temperature Tg, which may depend on the
applied pressure.[18][22] If the first-order freezing transition occurs over a range of
temperatures, and Tg falls within this range, then there is an interesting
possibility that the transition is arrested when it is partial and incomplete.
Extending these ideas to first-order magnetic transitions being arrested at low
temperatures, resulted in the observation of incomplete magnetic transitions,
with two magnetic phases coexisting, down to the lowest temperature. First
reported in the case of a ferromagnetic to anti-ferromagnetic transition, [23] such
persistent phase coexistence has now been reported across a variety of first-
order magnetic transitions. These include colossal-magnetoresistance manganite
materials,[24][25] magnetocaloric materials,[26] magnetic shape memory materials,
[27]
and other materials.[28] The interesting feature of these observations of Tg
falling within the temperature range over which the transition occurs is that the
first-order magnetic transition is influenced by magnetic field, just like the
structural transition is influenced by pressure. The relative ease with which
magnetic fields can be controlled, in contrast to pressure, raises the possibility
that one can study the interplay between Tg and Tc in an exhaustive way. Phase
coexistence across first-order magnetic transitions will then enable the resolution
of outstanding issues in understanding glasses.
Critical points[edit]
In any system containing liquid and gaseous phases, there exists a special
combination of pressure and temperature, known as the critical point, at which
the transition between liquid and gas becomes a second-order transition. Near
the critical point, the fluid is sufficiently hot and compressed that the distinction
between the liquid and gaseous phases is almost non-existent. This is associated
with the phenomenon of critical opalescence, a milky appearance of the liquid
due to density fluctuations at all possible wavelengths (including those of visible
light).
Symmetry[edit]
Phase transitions often involve a symmetry breaking process. For instance, the
cooling of a fluid into a crystalline solid breaks continuous translation symmetry:
each point in the fluid has the same properties, but each point in a crystal does
not have the same properties (unless the points are chosen from the lattice
points of the crystal lattice). Typically, the high-temperature phase contains more
symmetries than the low-temperature phase due to spontaneous symmetry
breaking, with the exception of certain accidental symmetries (e.g. the formation
of heavy virtual particles, which only occurs at low temperatures).[29]
Order parameters[edit]
Relevance in cosmology[edit]
Continuous phase transitions are easier to study than first-order transitions due
to the absence of latent heat, and they have been discovered to have many
interesting properties. The phenomena associated with continuous phase
transitions are called critical phenomena, due to their association with critical
points.
𝐶∝|𝑇c−𝑇|−𝛼.
The heat capacity of amorphous materials has such a behaviour near the glass
transition temperature where the universal critical exponent α = 0.59[33] A similar
behavior, but with the exponent ν instead of α, applies for the correlation length.
The exponent ν is positive. This is different with α. Its actual value depends on
the type of phase transition we are considering.
The critical exponents are not necessarily the same above and below the critical
temperature. When a continuous symmetry is explicitly broken down to a discrete
symmetry by irrelevant (in the renormalization group sense) anisotropies, then
some exponents (such as
For −1 < α < 0, the heat capacity has a "kink" at the transition temperature. This is
the behavior of liquid helium at the lambda transition from a normal state to the
superfluid state, for which experiments have found α = −0.013 ± 0.003. At least one
experiment was performed in the zero-gravity conditions of an orbiting satellite to
minimize pressure differences in the sample.[35] This experimental value of α agrees
with theoretical predictions based on variational perturbation theory.[36]
For 0 < α < 1, the heat capacity diverges at the transition temperature (though,
since α < 1, the enthalpy stays finite). An example of such behavior is the 3D
ferromagnetic phase transition. In the three-dimensional Ising model for uniaxial
magnets, detailed theoretical studies have yielded the exponent α ≈ +0.110.
Some model systems do not obey a power-law behavior. For example, mean field
theory predicts a finite discontinuity of the heat capacity at the transition
temperature, and the two-dimensional Ising model has a logarithmic divergence.
However, these systems are limiting cases and an exception to the rule. Real
phase transitions exhibit power-law behavior.
𝛽=𝛾/(𝛿−1),𝜈=𝛾/(2−𝜂).
It can be shown that there are only two independent exponents, e.g. ν and η.
Critical phenomena[edit]
There are also other critical phenomena; e.g., besides static functions there is
also critical dynamics. As a consequence, at a phase transition one may observe
critical slowing down or speeding up. Connected to the previous phenomenon is
also the phenomenon of enhanced fluctuations before the phase transition, as a
consequence of lower degree of stability of the initial phase of the system. The
large static universality classes of a continuous phase transition split into smaller
dynamic universality classes. In addition to the critical exponents, there are also
universal relations for certain static or dynamic functions of the magnetic fields
and temperature differences from the critical value.[citation needed]
It has been proposed that some biological systems might lie near critical points.
Examples include neural networks in the salamander retina,[40] bird flocks[41] gene
expression networks in Drosophila,[42] and protein folding.[43] However, it is not
clear whether or not alternative reasons could explain some of the phenomena
supporting arguments for criticality.[44] It has also been suggested that biological
organisms share two key properties of phase transitions: the change of
macroscopic behavior and the coherence of a system at a critical point. [45] Phase
transitions are prominent feature of motor behavior in biological systems. [46]
Spontaneous gait transitions,[47] as well as fatigue-induced motor task
disengagements,[48] show typical critical behavior as an intimation of the sudden
qualitative change of the previously stable motor behavioral pattern.
Experimental[edit]
A variety of methods are applied for studying the various effects. Selected
examples are:
See also[edit]
References[edit]
● ^ Askeland, Donald R.; Haddleton, Frank; Green, Phil; Robertson, Howard (1996). The
Science and Engineering of Materials. Chapman & Hall. p. 286. ISBN 978-0-412-53910-7.
● ^ Rybin, M.V.; et al. (2015). "Phase diagram for the transition from photonic crystals to
dielectric metamaterials". Nature Communications. 6: 10102. arXiv:1507.08901.
Bibcode:2015NatCo...610102R. doi:10.1038/ncomms10102. PMC 4686770. PMID
26626302.
● ^ Eds. Zhou, W., and Fan. S., Semiconductors and Semimetals. Vol 100. Photonic Crystal
Metasurface Optoelectronics, Elsevier, 2019
● ^ Carol Kendall (2004). "Fundamentals of Stable Isotope Geochemistry". USGS.
Retrieved 10 April 2014.
● ^
● Jump up to:
ab
● Jaeger, Gregg (1 May 1998). "The Ehrenfest Classification of Phase Transitions:
Introduction and Evolution". Archive for History of Exact Sciences. 53 (1): 51–81.
doi:10.1007/s004070050021. S2CID 121525126.
● ^
● Jump up to:
ab
● Blundell, Stephen J.; Katherine M. Blundell (2008). Concepts in Thermal Physics.
Oxford University Press. ISBN 978-0-19-856770-7.
● ^ Gross, David J. (1980), "Possible third-order phase transition in the large N lattice
gauge theory", Physical Review D, 21 (2): 446–453, doi:10.1103/PhysRevD.21.446
● ^ Majumdar, Satya N; Schehr, Grégory (31 January 2014). "Top eigenvalue of a random
matrix: large deviations and third order phase transition". Journal of Statistical
Mechanics: Theory and Experiment. 2014 (1): P01012. arXiv:1311.0580.
Bibcode:2014JSMTE..01..012M. doi:10.1088/1742-5468/2014/01/P01012. ISSN 1742-5468.
S2CID 119122520.
● ^
● Jump up to:
ab
● Pippard, Alfred B. (1981). Elements of classical thermodynamics: for advanced
students of physics (Repr ed.). Cambridge: Univ. Pr. pp. 140–141. ISBN 978-0-521-09101-
5.
● ^ Austin, J. B. (November 1932). "Heat Capacity of Iron - A Review". Industrial &
Engineering Chemistry. 24 (11): 1225–1235. doi:10.1021/ie50275a006. ISSN 0019-7866.
● ^ Jaeger, Gregg (1 May 1998). "The Ehrenfest Classification of Phase Transitions:
Introduction and Evolution". Archive for History of Exact Sciences. 53 (1): 51–81.
doi:10.1007/s004070050021. ISSN 1432-0657.
● ^ Stanley, H. Eugene (1971). Introduction to Phase Transitions and Critical Phenomena.
Oxford: Clarendon Press.
● ^ Faghri, A., and Zhang, Y., Transport Phenomena in Multiphase Systems, Elsevier,
Burlington, MA, 2006,
● ^ Faghri, A., and Zhang, Y., Fundamentals of Multiphase Heat Transfer and Flow,
Springer, New York, NY, 2020
● ^ Imry, Y.; Wortis, M. (1979). "Influence of quenched impurities on first-order phase
transitions". Phys. Rev. B. 19 (7): 3580–3585. Bibcode:1979PhRvB..19.3580I.
doi:10.1103/physrevb.19.3580.
● ^ Kumar, Kranti; Pramanik, A. K.; Banerjee, A.; Chaddah, P.; Roy, S. B.; Park, S.; Zhang,
C. L.; Cheong, S.-W. (2006). "Relating supercooling and glass-like arrest of kinetics for
phase separated systems: DopedCeFe2and(La,Pr,Ca)MnO3". Physical Review B. 73 (18):
184435. arXiv:cond-mat/0602627. Bibcode:2006PhRvB..73r4435K.
doi:10.1103/PhysRevB.73.184435. ISSN 1098-0121. S2CID 117080049.
● ^ Pasquini, G.; Daroca, D. Pérez; Chiliotte, C.; Lozano, G. S.; Bekeris, V. (2008).
"Ordered, Disordered, and Coexistent Stable Vortex Lattices inNbSe2Single Crystals".
Physical Review Letters. 100 (24): 247003. arXiv:0803.0307.
Bibcode:2008PhRvL.100x7003P. doi:10.1103/PhysRevLett.100.247003. ISSN 0031-9007.
PMID 18643617. S2CID 1568288.
● ^
● Jump up to:
ab
● Ojovan, M.I. (2013). "Ordering and structural changes at the glass-liquid transition".
J. Non-Cryst. Solids. 382: 79–86. Bibcode:2013JNCS..382...79O.
doi:10.1016/j.jnoncrysol.2013.10.016.
● ^ Gotze, Wolfgang. "Complex Dynamics of Glass-Forming Liquids: A Mode-Coupling
Theory."
● ^ Lubchenko, V. Wolynes; Wolynes, Peter G. (2007). "Theory of Structural Glasses and
Supercooled Liquids". Annual Review of Physical Chemistry. 58: 235–266. arXiv:cond-
mat/0607349. Bibcode:2007ARPC...58..235L.
doi:10.1146/annurev.physchem.58.032806.104653. PMID 17067282. S2CID 46089564.
● ^ Greer, A. L. (1995). "Metallic Glasses". Science. 267 (5206): 1947–1953.
Bibcode:1995Sci...267.1947G. doi:10.1126/science.267.5206.1947. PMID 17770105. S2CID
220105648.
● ^ Tarjus, G. (2007). "Materials science: Metal turned to glass". Nature. 448 (7155): 758–
759. Bibcode:2007Natur.448..758T. doi:10.1038/448758a. PMID 17700684. S2CID 4410586.
● ^ Manekar, M. A.; Chaudhary, S.; Chattopadhyay, M. K.; Singh, K. J.; Roy, S. B.;
Chaddah, P. (2001). "First-order transition from antiferromagnetism to ferromagnetism
inCe(Fe0.96Al0.04)2". Physical Review B. 64 (10): 104416. arXiv:cond-mat/0012472.
Bibcode:2001PhRvB..64j4416M. doi:10.1103/PhysRevB.64.104416. ISSN 0163-1829.
S2CID 16851501.
● ^ Banerjee, A.; Pramanik, A. K.; Kumar, Kranti; Chaddah, P. (2006). "Coexisting tunable
fractions of glassy and equilibrium long-range-order phases in manganites". Journal of
Physics: Condensed Matter. 18 (49): L605. arXiv:cond-mat/0611152.
Bibcode:2006JPCM...18L.605B. doi:10.1088/0953-8984/18/49/L02. S2CID 98145553.
● ^ Wu W.; Israel C.; Hur N.; Park S.; Cheong S. W.; de Lozanne A. (2006). "Magnetic
imaging of a supercooling glass transition in a weakly disordered ferromagnet". Nature
Materials. 5 (11): 881–886. Bibcode:2006NatMa...5..881W. doi:10.1038/nmat1743. PMID
17028576. S2CID 9036412.
● ^ Roy, S. B.; Chattopadhyay, M. K.; Chaddah, P.; Moore, J. D.; Perkins, G. K.; Cohen, L.
F.; Gschneidner, K. A.; Pecharsky, V. K. (2006). "Evidence of a magnetic glass state in
the magnetocaloric material Gd5Ge4". Physical Review B. 74 (1): 012403.
Bibcode:2006PhRvB..74a2403R. doi:10.1103/PhysRevB.74.012403. ISSN 1098-0121.
● ^ Lakhani, Archana; Banerjee, A.; Chaddah, P.; Chen, X.; Ramanujan, R. V. (2012).
"Magnetic glass in shape memory alloy: Ni45Co5Mn38Sn12". Journal of Physics:
Condensed Matter. 24 (38): 386004. arXiv:1206.2024. Bibcode:2012JPCM...24L6004L.
doi:10.1088/0953-8984/24/38/386004. ISSN 0953-8984. PMID 22927562. S2CID 206037831.
● ^ Kushwaha, Pallavi; Lakhani, Archana; Rawat, R.; Chaddah, P. (2009). "Low-
temperature study of field-induced antiferromagnetic-ferromagnetic transition in Pd-
doped Fe-Rh". Physical Review B. 80 (17): 174413. arXiv:0911.4552.
Bibcode:2009PhRvB..80q4413K. doi:10.1103/PhysRevB.80.174413. ISSN 1098-0121.
S2CID 119165221.
● ^ Ivancevic, Vladimir G.; Ivancevic, Tijiana, T. (2008). Complex Nonlinearity. Berlin:
Springer. pp. 176–177. ISBN 978-3-540-79357-1. Retrieved 12 October 2014.
● ^ Clark, J.B.; Hastie, J.W.; Kihlborg, L.H.E.; Metselaar, R.; Thackeray, M.M. (1994).
"Definitions of terms relating to phase transitions of the solid state". Pure and Applied
Chemistry. 66 (3): 577–594. doi:10.1351/pac199466030577. S2CID 95616565.
● ^ Chaisson, Eric J. (2001). Cosmic Evolution. Harvard University Press. ISBN 978-0-674-
00342-2.
● ^ David Layzer, Cosmogenesis, The Development of Order in the Universe, Oxford Univ.
Press, 1991
● ^ Ojovan, Michael I.; Lee, William E. (2006). "Topologically disordered systems at the
glass transition" (PDF). Journal of Physics: Condensed Matter. 18 (50): 11507–11520.
Bibcode:2006JPCM...1811507O. doi:10.1088/0953-8984/18/50/007. S2CID 96326822.
● ^ Leonard, F.; Delamotte, B. (2015). "Critical exponents can be different on the two sides
of a transition". Phys. Rev. Lett. 115 (20): 200601. arXiv:1508.07852.
Bibcode:2015PhRvL.115t0601L. doi:10.1103/PhysRevLett.115.200601. PMID 26613426.
S2CID 22181730.
● ^ Lipa, J.; Nissen, J.; Stricker, D.; Swanson, D.; Chui, T. (2003). "Specific heat of liquid
helium in zero gravity very near the lambda point". Physical Review B. 68 (17): 174518.
arXiv:cond-mat/0310163. Bibcode:2003PhRvB..68q4518L.
doi:10.1103/PhysRevB.68.174518. S2CID 55646571.
● ^ Kleinert, Hagen (1999). "Critical exponents from seven-loop strong-coupling φ4 theory
in three dimensions". Physical Review D. 60 (8): 085001. arXiv:hep-th/9812197.
Bibcode:1999PhRvD..60h5001K. doi:10.1103/PhysRevD.60.085001. S2CID 117436273.
● ^ D.Y. Lando and V.B. Teif (2000). "Long-range interactions between ligands bound to a
DNA molecule give rise to adsorption with the character of phase transition of the first
kind". J. Biomol. Struct. Dyn. 17 (5): 903–911. doi:10.1080/07391102.2000.10506578. PMID
10798534. S2CID 23837885.
13
● ^ Yashroy RC (1987). " C NMR studies of lipid fatty acyl chains of chloroplast
membranes". Indian Journal of Biochemistry and Biophysics. 24 (6): 177–178.
doi:10.1016/0165-022X(91)90019-S. PMID 3428918.
● ^ YashRoy, R C (1990). "Determination of membrane lipid phase transition temperature
from 13-C NMR intensities". Journal of Biochemical and Biophysical Methods. 20 (4):
353–356. doi:10.1016/0165-022X(90)90097-V. PMID 2365951.
● ^ Tkacik, Gasper; Mora, Thierry; Marre, Olivier; Amodei, Dario; Berry II, Michael J.;
Bialek, William (2014). "Thermodynamics for a network of neurons: Signatures of
criticality". arXiv:1407.5946 [q-bio.NC].
● ^ Bialek, W; Cavagna, A; Giardina, I (2014). "Social interactions dominate speed control
in poising natural flocks near criticality". PNAS. 111 (20): 7212–7217. arXiv:1307.5563.
Bibcode:2014PNAS..111.7212B. doi:10.1073/pnas.1324045111. PMC 4034227. PMID
24785504.
● ^ Krotov, D; Dubuis, J O; Gregor, T; Bialek, W (2014). "Morphogenesis at criticality".
PNAS. 111 (10): 3683–3688. arXiv:1309.2614. Bibcode:2014PNAS..111.3683K.
doi:10.1073/pnas.1324186111. PMC 3956198. PMID 24516161.
● ^ Mora, Thierry; Bialek, William (2011). "Are biological systems poised at criticality?".
Journal of Statistical Physics. 144 (2): 268–302. arXiv:1012.2242.
Bibcode:2011JSP...144..268M. doi:10.1007/s10955-011-0229-4. S2CID 703231.
● ^ Schwab, David J; Nemenman, Ilya; Mehta, Pankaj (2014). "Zipf's law and criticality in
multivariate data without fine-tuning". Physical Review Letters. 113 (6): 068102.
arXiv:1310.0448. Bibcode:2014PhRvL.113f8102S. doi:10.1103/PhysRevLett.113.068102.
PMC 5142845. PMID 25148352.
● ^ Longo, G.; Montévil, M. (1 August 2011). "From physics to biology by extending
criticality and symmetry breakings". Progress in Biophysics and Molecular Biology.
Systems Biology and Cancer. 106 (2): 340–347. arXiv:1103.1833.
doi:10.1016/j.pbiomolbio.2011.03.005. PMID 21419157. S2CID 723820.
● ^ Kelso, J. A. Scott (1995). Dynamic Patterns: The Self-Organization of Brain and
Behavior (Complex Adaptive Systems). MIT Press. ISBN 978-0-262-61131-2.
● ^ Diedrich, F. J.; Warren, W. H. Jr. (1995). "Why change gaits? Dynamics of the walk-run
transition". Journal of Experimental Psychology. Human Perception and Performance.
21 (1): 183–202. doi:10.1037/0096-1523.21.1.183. PMID 7707029.
● ^ Hristovski, R.; Balagué, N. (2010). "Fatigue-induced spontaneous termination point--
nonequilibrium phase transitions and critical behavior in quasi-isometric exertion".
Human Movement Science. 29 (4): 483–493. doi:10.1016/j.humov.2010.05.004. PMID
20619908.
● ^ Moret, Marcelo; Zebende, Gilney (January 2007). "Amino acid hydrophobicity and
accessible surface area". Physical Review E. 75 (1): 011920.
Bibcode:2007PhRvE..75a1920M. doi:10.1103/PhysRevE.75.011920. PMID 17358197.
● ^ Gorban, A.N.; Smirnova, E.V.; Tyukina, T.A. (August 2010). "Correlations, risk and
crisis: From physiology to finance". Physica A: Statistical Mechanics and Its
Applications. 389 (16): 3193–3217. arXiv:0905.0129. Bibcode:2010PhyA..389.3193G.
doi:10.1016/j.physa.2010.03.035. S2CID 276956.
Further reading[edit]
External links[edit]
Wikimedia Commons has media related to Phase changes.
show
● V
● T
● E
Authority control
● France
databases: National
● BnF data
● Germany
● Israel
● United States
● Japan
● Czech Republic
Categories:
Phase transitions
Physical phenomena
Critical phenomena
This page was last edited on 21 May 2024, at 06:31 (UTC).
Text is available under the Creative Commons Attribution-ShareAlike License 4.0; additional terms may
apply. By using this site, you agree to the Terms of Use and Privacy Policy. Wikipedia® is a
registered trademark of the Wikimedia Foundation, Inc., a non-profit or