You are on page 1of 51

Melting point

Ice cubes put in water will reach the 0 °C melting point


of ice when they start to melt.
The melting point (or, rarely, liquefaction
point) of a substance is the temperature at
which it changes state from solid to liquid.
At the melting point the solid and liquid
phase exist in equilibrium. The melting
point of a substance depends on pressure
and is usually specified at a standard
pressure such as 1 atmosphere or 100
kPa.

When considered as the temperature of


the reverse change from liquid to solid, it is
referred to as the freezing point or
crystallization point. Because of the ability
of some substances to supercool, the
freezing point is not considered as a
characteristic property of a substance.
When the "characteristic freezing point" of
a substance is determined, in fact the
actual methodology is almost always "the
principle of observing the disappearance
rather than the formation of ice, that is, the
melting point.[1]

Examples
Melting points (in blue) and boiling points (in pink) of
the first eight carboxylic acids (°C)

For most substances, melting and freezing


points are approximately equal. For
example, the melting point and freezing
point of mercury is 234.32 Kelvin
(−38.83 °C or −37.89 °F).[2] However,
certain substances possess differing
solid-liquid transition temperatures. For
example, agar melts at 85 °C (185 °F) and
solidifies from 31 °C (88 °F; 304 K); such
direction dependence is known as
hysteresis. The melting point of ice at 1
atmosphere of pressure is very close [3] to
0 °C (32 °F; 273 K); this is also known as
the ice point. In the presence of nucleating
substances, the freezing point of water is
not always the same as the melting point.
In the absence of nucleators water can
exist as a supercooled liquid down to
−48.3 °C (−55 °F, 224.8 K) before freezing.
The chemical element with the highest
melting point is tungsten, at 3,414 °C
(6,177 °F; 3,687 K);[4] this property makes
tungsten excellent for use as filaments in
light bulbs. The often-cited carbon does
not melt at ambient pressure but sublimes
at about 3,726.85 °C (6,740.33 °F;
4,000.00 K); a liquid phase only exists
above pressures of 10 MPa (99 atm) and
estimated 4,030–4,430 °C (7,290–
8,010 °F; 4,300–4,700 K) (see carbon
phase diagram). Tantalum hafnium
carbide (Ta4HfC5) is a refractory
compound with a very high melting point
of 4215 K (3942 °C, 7128 °F).[5] Quantum
mechanical computer simulations have
predicted that the alloy HfN0.38C0.51 will
have an even higher melting point (about
4400 K),[6] which would make it the
substance with the highest melting point
at ambient pressure. This prediction was
later confirmed by experiment.[7] At the
other end of the scale, helium does not
freeze at all at normal pressure even at
temperatures arbitrarily close to absolute
zero; a pressure of more than twenty times
normal atmospheric pressure is
necessary.
List of common chemicals
g
Chemical[I] Density ( cm3 ) Melt (K) [8] Boil (K)

Water @STP 1 273 373

Solder (Pb60Sn40) 456

Cocoa butter 307.2 -

Paraffin wax 0.9 310 643

Hydrogen 0.00008988 14.01 20.28

Helium 0.0001785 —[II] 4.22

Beryllium 1.85 1560 2742

Carbon 2.267 —[III][9] 4000[III][9]

Nitrogen 0.0012506 63.15 77.36

Oxygen 0.001429 54.36 90.20

Sodium 0.971 370.87 1156

Magnesium 1.738 923 1363

Aluminium 2.698 933.47 2792

Sulfur 2.067 388.36 717.87

Chlorine 0.003214 171.6 239.11

Potassium 0.862 336.53 1032

Titanium 4.54 1941 3560

Iron 7.874 1811 3134

Nickel 8.912 1728 3186

Copper 8.96 1357.77 2835

Zinc 7.134 692.88 1180

Gallium 5.907 302.9146 2673


Silver 10.501 1234.93 2435

Cadmium 8.69 594.22 1040

Indium 7.31 429.75 2345

Iodine 4.93 386.85 457.4

Tantalum 16.654 3290 5731

Tungsten 19.25 3695 5828

Platinum 21.46 2041.4 4098

Gold 19.282 1337.33 3129

Mercury 13.5336 234.43 629.88

Lead 11.342 600.61 2022

Bismuth 9.807 544.7 1837

Notes

I. Z is the standard symbol for atomic number; C is the standard symbol for heat capacity; and
χ is the standard symbol for electronegativity on the Pauling scale.

II. Helium does not solidify at a pressure of one atmosphere. Helium can only solidify at
pressures above 25 atmospheres, which corresponds to a melting point of absolute zero.

III. Carbon does not melt at any temperature under standard pressure, instead it sublimes
around 4100K

Melting point measurements


Kofler bench with samples for calibration

Many laboratory techniques exist for the


determination of melting points. A Kofler
bench is a metal strip with a temperature
gradient (range from room temperature to
300 °C). Any substance can be placed on a
section of the strip, revealing its thermal
behaviour at the temperature at that point.
Differential scanning calorimetry gives
information on melting point together with
its enthalpy of fusion.

Automatic digital melting point meter

A basic melting point apparatus for the


analysis of crystalline solids consists of
an oil bath with a transparent window
(most basic design: a Thiele tube) and a
simple magnifier. Several grains of a solid
are placed in a thin glass tube and partially
immersed in the oil bath. The oil bath is
heated (and stirred) and with the aid of the
magnifier (and external light source)
melting of the individual crystals at a
certain temperature can be observed. A
metal block might be used instead of an
oil bath. Some modern instruments have
automatic optical detection.

The measurement can also be made


continuously with an operating process.
For instance, oil refineries measure the
freeze point of diesel fuel "online",
meaning that the sample is taken from the
process and measured automatically. This
allows for more frequent measurements
as the sample does not have to be
manually collected and taken to a remote
laboratory.

Techniques for refractory materials …

For refractory materials (e.g. platinum,


tungsten, tantalum, some carbides and
nitrides, etc.) the extremely high melting
point (typically considered to be above,
say, 1800 °C) may be determined by
heating the material in a black body
furnace and measuring the black-body
temperature with an optical pyrometer. For
the highest melting materials, this may
require extrapolation by several hundred
degrees. The spectral radiance from an
incandescent body is known to be a
function of its temperature. An optical
pyrometer matches the radiance of a body
under study to the radiance of a source
that has been previously calibrated as a
function of temperature. In this way, the
measurement of the absolute magnitude
of the intensity of radiation is
unnecessary. However, known
temperatures must be used to determine
the calibration of the pyrometer. For
temperatures above the calibration range
of the source, an extrapolation technique
must be employed. This extrapolation is
accomplished by using Planck's law of
radiation. The constants in this equation
are not known with sufficient accuracy,
causing errors in the extrapolation to
become larger at higher temperatures.
However, standard techniques have been
developed to perform this extrapolation.

Consider the case of using gold as the


source (mp = 1063 °C). In this technique,
the current through the filament of the
pyrometer is adjusted until the light
intensity of the filament matches that of a
black-body at the melting point of gold.
This establishes the primary calibration
temperature and can be expressed in
terms of current through the pyrometer
lamp. With the same current setting, the
pyrometer is sighted on another black-
body at a higher temperature. An
absorbing medium of known transmission
is inserted between the pyrometer and this
black-body. The temperature of the black-
body is then adjusted until a match exists
between its intensity and that of the
pyrometer filament. The true higher
temperature of the black-body is then
determined from Planck's Law. The
absorbing medium is then removed and
the current through the filament is
adjusted to match the filament intensity to
that of the black-body. This establishes a
second calibration point for the pyrometer.
This step is repeated to carry the
calibration to higher temperatures. Now,
temperatures and their corresponding
pyrometer filament currents are known
and a curve of temperature versus current
can be drawn. This curve can then be
extrapolated to very high temperatures.
In determining melting points of a
refractory substance by this method, it is
necessary to either have black body
conditions or to know the emissivity of the
material being measured. The
containment of the high melting material
in the liquid state may introduce
experimental difficulties. Melting
temperatures of some refractory metals
have thus been measured by observing the
radiation from a black body cavity in solid
metal specimens that were much longer
than they were wide. To form such a cavity,
a hole is drilled perpendicular to the long
axis at the center of a rod of the material.
These rods are then heated by passing a
very large current through them, and the
radiation emitted from the hole is
observed with an optical pyrometer. The
point of melting is indicated by the
darkening of the hole when the liquid
phase appears, destroying the black body
conditions. Today, containerless laser
heating techniques, combined with fast
pyrometers and spectro-pyrometers, are
employed to allow for precise control of
the time for which the sample is kept at
extreme temperatures. Such experiments
of sub-second duration address several of
the challenges associated with more
traditional melting point measurements
made at very high temperatures, such as
sample vaporization and reaction with the
container.

Thermodynamics
Pressure dependence of water melting point.

For a solid to melt, heat is required to raise


its temperature to the melting point.
However, further heat needs to be supplied
for the melting to take place: this is called
the heat of fusion, and is an example of
latent heat.
From a thermodynamics point of view, at
the melting point the change in Gibbs free
energy (ΔG) of the material is zero, but the
enthalpy (H) and the entropy (S) of the
material are increasing (ΔH, ΔS > 0).
Melting phenomenon happens when the
Gibbs free energy of the liquid becomes
lower than the solid for that material. At
various pressures this happens at a
specific temperature. It can also be shown
that:
Here T, ΔS and ΔH are respectively the
temperature at the melting point, change
of entropy of melting and the change of
enthalpy of melting.

The melting point is sensitive to extremely


large changes in pressure, but generally
this sensitivity is orders of magnitude less
than that for the boiling point, because the
solid-liquid transition represents only a
small change in volume.[10][11] If, as
observed in most cases, a substance is
more dense in the solid than in the liquid
state, the melting point will increase with
increases in pressure. Otherwise the
reverse behavior occurs. Notably, this is
the case of water, as illustrated graphically
to the right, but also of Si, Ge, Ga, Bi. With
extremely large changes in pressure,
substantial changes to the melting point
are observed. For example, the melting
point of silicon at ambient pressure (0.1
MPa) is 1415 °C, but at pressures in
excess of 10 GPa it decreases to
1000 °C.[12]
Melting points are often used to
characterize organic and inorganic
compounds and to ascertain their purity.
The melting point of a pure substance is
always higher and has a smaller range
than the melting point of an impure
substance or, more generally, of mixtures.
The higher the quantity of other
components, the lower the melting point
and the broader will be the melting point
range, often referred to as the "pasty
range". The temperature at which melting
begins for a mixture is known as the
"solidus" while the temperature where
melting is complete is called the "liquidus".
Eutectics are special types of mixtures
that behave like single phases. They melt
sharply at a constant temperature to form
a liquid of the same composition.
Alternatively, on cooling a liquid with the
eutectic composition will solidify as
uniformly dispersed, small (fine-grained)
mixed crystals with the same composition.

In contrast to crystalline solids, glasses do


not possess a melting point; on heating
they undergo a smooth glass transition
into a viscous liquid. Upon further heating,
they gradually soften, which can be
characterized by certain softening points.

Freezing-point depression
The freezing point of a solvent is
depressed when another compound is
added, meaning that a solution has a
lower freezing point than a pure solvent.
This phenomenon is used in technical
applications to avoid freezing, for instance
by adding salt or ethylene glycol to water.

Carnelley's rule
In organic chemistry, Carnelley's rule,
established in 1882 by Thomas Carnelley,
states that high molecular symmetry is
associated with high melting point.[13]
Carnelley based his rule on examination of
15,000 chemical compounds. For
example, for three structural isomers with
molecular formula C5H12 the melting point
increases in the series isopentane −160 °C
(113 K) n-pentane −129.8 °C (143 K) and
neopentane −16.4 °C (256.8 K).[14]
Likewise in xylenes and also
dichlorobenzenes the melting point
increases in the order meta, ortho and
then para. Pyridine has a lower symmetry
than benzene hence its lower melting
point but the melting point again increases
with diazine and triazines. Many cage-like
compounds like adamantane and cubane
with high symmetry have relatively high
melting points.
A high melting point results from a high
heat of fusion, a low entropy of fusion, or a
combination of both. In highly symmetrical
molecules the crystal phase is densely
packed with many efficient intermolecular
interactions resulting in a higher enthalpy
change on melting.

Like many high symmetry compounds,


tetrakis(trimethylsilyl)silane has a very high melting
point (m.p.) of 319-321 °C. It tends to sublime, so the
m.p. determination requires that the sample be sealed
p q p
in a tube.[15]

Predicting the melting point


of substances (Lindemann's
criterion)
An attempt to predict the bulk melting
point of crystalline materials was first
made in 1910 by Frederick Lindemann.[16]
The idea behind the theory was the
observation that the average amplitude of
thermal vibrations increases with
increasing temperature. Melting initiates
when the amplitude of vibration becomes
large enough for adjacent atoms to partly
occupy the same space. The Lindemann
criterion states that melting is expected
when the vibration root mean square
amplitude exceeds a threshold value.

Assuming that all atoms in a crystal


vibrate with the same frequency ν, the
average thermal energy can be estimated
using the equipartition theorem as[17]
where m is the atomic mass, ν is the
frequency, u is the average vibration
amplitude, kB is the Boltzmann constant,
and T is the absolute temperature. If the
threshold value of u2 is c2a2 where c is the
Lindemann constant and a is the atomic
spacing, then the melting point is
estimated as

Several other expressions for the


estimated melting temperature can be
obtained depending on the estimate of the
average thermal energy. Another
commonly used expression for the
Lindemann criterion is[18]

From the expression for the Debye


frequency for ν, we have
where θD is the Debye temperature and h
is the Planck constant. Values of c range
from 0.15–0.3 for most materials.[19]

Melting point prediction


In February 2011, Alfa Aesar released over
10,000 melting points of compounds from
their catalog as open data. This dataset
has been used to create a random forest
model for melting point prediction which is
now freely available.[20] Open melting point
data are also available from Nature
Precedings.[21] High quality data mined
from patents and also models[22]
developed with these data were published
by Tetko et al.[23]

See also
Heat
Highest melting point
Liquidus
List of elements by melting point
Melting
Melting points of the elements (data
page)
Phase diagram
Phases of matter
Pressure
Simon–Glatzel equation
Slip melting point
Solidus temperature
Triple point
Zone melting

References
Citations …

1. Ramsay, J. A. (1 May 1949). "A New


Method of Freezing-Point
Determination for Small Quantities" .
Journal of Experimental Biology. 26
(1): 57–64. PMID 15406812 .
2. Haynes, p. 4.122.
3. The melting point of purified water has
been measured as 0.002519 ±
0.000002 °C, see Feistel, R. & Wagner,
W. (2006). "A New Equation of State
for H2O Ice Ih". J. Phys. Chem. Ref.
Data. 35 (2): 1021–1047.
Bibcode:2006JPCRD..35.1021F .
doi:10.1063/1.2183324 .
4. Haynes, p. 4.123.
5. Agte, C. & Alterthum, H. (1930).
"Researches on Systems with
Carbides at High Melting Point and
Contributions to the Problem of
Carbon Fusion". Z. Tech. Phys. 11:
182–191.
. Hong, Q.-J.; van de Walle, A. (2015).
"Prediction of the material with highest
known melting point from ab initio
molecular dynamics calculations".
Phys. Rev. B. 92 (2): 020104(R).
Bibcode:2015PhRvB..92b0104H .
doi:10.1103/PhysRevB.92.020104 .
7. Buinevich, V.S.; Nepapushev, A.A.;
Moskovskikh, D.O.; Trusov, G.V.;
Kuskov, K.V.; Vadchenko, S.G.;
Rogachev, A.S.; Mukasyan, A.S. (March
2020). "Fabrication of ultra-high-
temperature nonstoichiometric
hafnium carbonitride via combustion
synthesis and spark plasma sintering".
Ceramics International.
doi:10.1016/j.ceramint.2020.03.158 .
. Holman, S. W.; Lawrence, R. R.; Barr, L.
(1 January 1895). "Melting Points of
Aluminum, Silver, Gold, Copper, and
Platinum". Proceedings of the
American Academy of Arts and
Sciences. 31: 218–233.
doi:10.2307/20020628 .
JSTOR 20020628 .
9. "Carbon" . rsc.org.
10. The exact relationship is expressed in
the Clausius–Clapeyron relation.
11. "J10 Heat: Change of aggregate state
of substances through change of heat
content: Change of aggregate state of
substances and the equation of
Clapeyron-Clausius" . Retrieved
19 February 2008.
12. Tonkov, E. Yu. and Ponyatovsky, E. G.
(2005) Phase Transformations of
Elements Under High Pressure, CRC
Press, Boca Raton, p. 98 ISBN 0-8493-
3367-9
13. Brown, R. J. C. & R. F. C. (2000).
"Melting Point and Molecular
Symmetry". Journal of Chemical
Education. 77 (6): 724.
Bibcode:2000JChEd..77..724B .
doi:10.1021/ed077p724 .
14. Haynes, pp. 6.153–155.
15. Gilman, H.; Smith, C. L. (1967).
"Tetrakis(trimethylsilyl)silane". Journal
of Organometallic Chemistry. 8 (2):
245–253. doi:10.1016/S0022-
328X(00)91037-4 .
1 . Lindemann FA (1910). "The calculation
of molecular vibration frequencies".
Phys. Z. 11: 609–612.
17. Sorkin, S., (2003), Point defects, lattice
structure, and melting , Thesis,
Technion, Israel.
1 . Philip Hofmann (2008). Solid state
physics: an introduction . Wiley-VCH.
p. 67. ISBN 978-3-527-40861-0.
Retrieved 13 March 2011.
19. Nelson, D. R., (2002), Defects and
geometry in condensed matter
physics , Cambridge University Press,
ISBN 0-521-00400-4
20. Predict melting point from SMILES .
Qsardb.org. Retrieved on 13
September 2013.
21. Bradley, Jean-Claude; Lang, Andrew;
Williams, Antony; Curtin, Evan (11
August 2011). "ONS Open Melting
Point Collection". Nature Precedings.
doi:10.1038/npre.2011.6229.1 .
22. OCHEM melting point models .
ochem.eu. Retrieved on 18 June 2016.
23. Tetko, Igor V; m. Lowe, Daniel;
Williams, Antony J (2016). "The
development of models to predict
melting and pyrolysis point data
associated with several hundred
thousand compounds mined from
PATENTS" . Journal of
Cheminformatics. 8: 2.
doi:10.1186/s13321-016-0113-y .
PMC 4724158 . PMID 26807157 .

Sources …
Works cited
Haynes, William M., ed. (2011). CRC
Handbook of Chemistry and Physics (92nd
ed.). CRC Press. ISBN 978-1439855119.

External links
Melting and boiling point tables vol. 1
by Thomas Carnelley (Harrison, London,
1885–1887)
Melting and boiling point tables vol. 2
by Thomas Carnelley (Harrison, London,
1885–1887)
Patent mined data Over 250,000 freely
downloadable melting point data. Also
downloadable at figshare

Retrieved from
"https://en.wikipedia.org/w/index.php?
title=Melting_point&oldid=956472211"

Last edited 8 days ago by Chris Capoccia

Content is available under CC BY-SA 3.0 unless


otherwise noted.

You might also like