You are on page 1of 10

The Forces Causing Phase Transition of Matter

Jerry Z. Liu

Stanford University, California, USA

Abstract

Phase transition of matter is a well known phenomenon and well studied subject.
However, what are the forces that cause the phase transition is not necessarily well
understood. An orbital repulsion force is investigated in this study. Upon absorbing
energy, an electron jumps to a higher orbital. The process not only increase the size of
an atom or molecule, but also reduces the distance to the adjacent molecules. Both
processes increase the repulsions to the adjacent molecules. Because an orbital jump is
spontaneous, the repulsions may exert a significant force to the adjacent molecules. This
force may play a critical role in phase transition. Intermolecular force is also model in this
study using two helium atoms to understand the trends in terms of atom size and
distance between atoms. With the new understanding, we are able to tell the nature
behind the melting/boiling point and supercritical point. They are all determined by the
orbital repulsion force and the intermolecular force.

Introduction

Phase transition of matter are phenomena observable every day. It has been observed that the
phenomena are related to temperature and pressure. It is further understood that these
phenomena are related to the intermolecular forces which in turn are affected by temperature
and pressure. This study is intended to look into the details of these interactions. In particular,
we will study a new type of intermolecular force, namely ​orbital repulsion force. It interacts with
other intermolecular force and plays a critical role in phase transition.

Before moving forward, we need to clarify some nomenclature related to the forces that are
interactions in the phase transition. Intermolecular force are the forces which mediate interaction
between molecules, including both attraction and repulsion forces. The forces are categorized
into hydrogen bonding, ionic bonding, ion-induced dipole forces, ion-dipole forces and van der
Waals forces.​[1] An intramolecular force is any force that binds together the atoms making up a
molecule or compound, such as covalent bonding and ionic bonding.​[2-3] By definition, these two
terms should refer two sets of disjointed forces. However, both claimed the ionic bonding, which
make it somewhat confusion. At low temperatures, ions are held together in molecules by ionic
bonds. However, at very high temperatures, such as in plasma, ions are separated. It is further
confused with van der Waals force, which is sometimes used loosely for all intermolecular
forces. The van der Waals force is a distance-dependent interaction between atoms or
molecules, usually including Keesom force, Debye force, and London dispersion force.​[4-6] To
avoid confusion in our discussions, the term intermolecular force is used to refer to the
definition: mediate interaction between molecules. Therefore, in a solid, an ionic bond is not
count as intermolecular force. All these named forces are electrostatic in nature, which is
described by Coulomb’s law:

F​ = ​Kq​1q ​ r​ ​-2,​
​ 2 (1)

where ​K is Coulomb’s constant, positive ​F is the repulsive force between signed charges ​q​1 and
q2​ ,​ and ​r is the distance between the charges.​[7-9] It is an electric analogy to Isaac Newton’s
universal gravitation.​[10-11] While gravity governs the interactions among celestial bodies in
classical mechanics, Coulomb’s force is most prominent at the subatomic realm because it is
10​36​ magnitude stronger than the gravitational force.​[12]

Intermolecular Force

To investigate intermolecular forces, we propose further scrutiny of the forces between


molecules. Normally, each molecule is electrically neutral. However, it does not mean there is
no electric force between molecules. Consider a model where electrons orbit around their
nucleus in a cloud and the charge density is evenly distributed in the shells. The repulsion
against a negatively charged particle outside the atom can be computed by integrating the force
over the shell. The resulted accumulative repulsion from the shells is canceled exactly by the
attraction of their nucleus, producing a zero net repulsion to any particle outside the atom. This
is not consistent with observations, because there are attractions holding molecules together in
a solid. Then, how can we model the force between molecules? To establish the model, we
need an accurate atom model in the first place. Unfortunately, a general solution to the
Schrödinger equation is not available other than the hydrogen atom.​[13] Due to the dynamic
nature of electrons, it is difficult to model intermolecular forces. Nevertheless, one thing we can
do at least is to find the bounding values for the repulsion or attraction forces using a simplified
model. Figure 1 shows two configurations for two helium atoms next to each other. The
maximum repulsion force is found in configuration 1A and minimum in 1B. In configuration 1B,
the two electrons orbital in the left atom is in a plane perpendicular to the centerline between the
atoms.
Figure 1, Configurations for helium atom repulsion model.

Based on Coulomb's law in equation (1), the net repulsion forces between the two helium atoms
for configuration 1A and 1B are:

Fa​ ​ = Kqe​ ​2​(s-2


​ ​ + 6(s+2r)​-2​ + (s+4r)​-2​ - 4(s+r)-2
​ ​ - 4(s+3r)-2​
​ ), (2)
and

F​b​ = Kq​e​2(4(s+2r)​-2​ - 2(s+r)-2
​ ​ - 2(s+3r)-2​ ​ + 2(s+r)(r​2+(r+s)

​2​)​-1.5​ +
2(s+3r)(r​2​+(s+3r)​2​)-1.5
​ ​ - 4(s+2r)(r​2+(s+2r)

​2​)​-1.5​), (3)

where ​F​a and ​F​b are the maximum and minimum repulsion forces, ​K is the Coulomb’s constant,
qe​ is the charge of an electron, ​r is the radius of the helium atoms and ​s is the distance between
the two atoms. Using the model, we can investigate the repulsion trends in terms of atom sizes
and distances between the atoms. For instance, given a helium atom radius 140 pm, Figure 2
shows the repulsion trends as a function of atom distances. The maximum repulsion decreases
quickly with increasing distances. It is interesting to note that the minimum repulsion is negative,
meaning there is an attraction between the two atoms in configuration 1B. The attraction
decreases with increasing distance. Both values decrease to zero at infinite distance. It is
important to understand that the two configurations are extreme cases. The repulsion force can
be upto 5.4 nano N at a distance of 140 pm (one atom radius) and the attraction force can be
upto 1.03 nano N. Since electrons are very dynamic particles, the actual configuration at any
moment is something in between. So, for a given distance, the actual force must be a value
between the two curves. Figure 2 also shows an important trend for the range between
maximum and minimum forces, which increases with decreasing distances between atoms.
With a small distance, the range is very large and may fluctuate from repulsion all the way to
attraction. In other words, the force can be either repulsion or attraction. It all depends on the
actual configuration of the electron clouds.
Figure 2, Maximum forces between two helium atoms as a function of distance.

It is also interesting to look into the repulsion trends as a function of atom sizes. The repulsion
increases with atom sizes for a fixed atom distance. Three curves, blue, red and yellow, are
plotted in Figure 3 for different distances, 20, 30 and 40 pm. In general, the repulsion increases
with atom sizes for all the three distances. When atom size increases, the distance between the
nucleus and electrons of adjacent atoms increases, which reduces the attractive components in
equation (2) of the model.

Figure 3, Repulsion as a function of atom sizes for three different atom distances.
Even though the above model is based on simple configurations with two helium atoms, the
results are generally applicable to most atoms, especially for large ones. Compared with helium,
a hydrogen atom has only one proton and exerts less attraction to its electron. So, its atom size
is much larger. The same model for hydrogen atoms is simpler and the general trends are
similar, but with larger maximum repulsion/attraction ranges. The force modeled using helium
and hydrogen represents the two extreme cases, where helium has fully filled in the outermost
shell while hydrogen is minimally filled. For other elements, the force will be in between of the
two models. With more electrons in large atoms, the repulsion between electrons causes them
to be distributed more evenly in their shells. As mentioned earlier the evenly distributed electron
clouds cancel the attraction from their nucleus to an outside electron. Hence, only the outermost
electrons are significant to the force between atoms. So, the force tends to be similar to either
helium model or hydrogen model depending on the structure of the outermost orbitals.

Upon absorption of energy, electrons jump to higher orbitals, which is effectively increased the
sizes of atoms or molecules. Based on the above model, the repulsion between atoms
increases with atom size. Hence, in general, the repulsion between molecules increases with
temperature. This explains the nature of thermal expansion.

Orbital Repulsion Force

Orbital repulsion force is a spontaneous short-range intermolecular force created in the process
of electron orbital jump.​[14] Upon absorption of energy or photons, an electron in an atom jumps
to a higher orbital, also known as orbital jump. The process of orbital jump is still under active
research. It appears spontaneous as an electron jumps from one energy level to another,
typically in nanoseconds or less. Because an orbital jump is sudden, it may exert a repulsion
force to the object next to the molecule. The object being pushed may be anything range from
another molecule, a dust particle, to a wall of the container. As shown in Figure 4A, an atom is
next to an object initially. On absorbing energy, the atom goes through an orbital jump process,
Figure 4B, which actually includes two sub-processes. First, it increases the atom size, which
increases the repulsion between the atom and the object. Second, the distance between them is
decreased, which also increases the repulsion further. Both processes contribute to the
repulsion force together in a process called ​orbital repulsion​.
Figure 4, Orbital repulsion resulted from orbital jump.

The force created in the orbital repulsion process may be significant at a short range. Recall, the
repulsion force between the helium atoms at a distance of 140 pm may be upto 5.4 nano N. To
get a sense of the scale of this force, that is equivalent to accelerating a kilogram mass with
8.12x10​17 N force. This is not a conventional intermolecular force, which occurs only when there
is orbital jump. This force may be responsible for Brownian motion and drive the Crookes
radiometer. The force may play an important role in phase transition of matter.

Molecular Gravity

The concept of van der Waals radius was developed to recognize that atoms were not
points.​[15-16] In fact, the van der Waals radius is estimated based on the volume of gas. The van
der Waals distance is defined as a distance between molecules that the intermolecular force
becomes repulsive when distance between molecules is less than the van der Waals distance.
Half of the van der Waals distance is defined as van der Waals radius. When the distance
between molecules is smaller than the van der Waals distance, the repulsion will push them
apart. Hence, it is the van der Waals radius that determine the volume of a system. The concept
may be generalized to liquid and solid without much problem, which explains the volume and
density of matter in general. The van der Waals distance is a balanced distance between
repulsion and attraction forces between molecules. In response to temperature change, both
molecule size and van der Waals distance must be adjusted to find a new balance, which the
which is the thermal expansion or attraction.
Figure 5, Molecular gravity due to the attraction between molecules from far side

However, these are not enough to explain phase transition. In most of the models for
intermolecular forces, only two molecules are considered, including our model. To understand
phase transition, multiple molecules must be considered collectively. Even though the attraction
and repulsion forces are balanced between adjacent molecules as shown by the red lines in
Figure 5, the forces between molecules at distances are attractive as shown in the figure by the
blue lines. Effect of these force is somewhat like gravitational force. So, let’s call it ​molecular
gravity​. The total molecular gravity for a given molecule is the pressure of all the molecules
surround it except the molecules right next to it. The force can be computed by integrating all
the attractive forces from surrounding molecules. The van der Waals force decreases with
distance between molecules. Because molecule size and distance between molecules increase
with temperature. The molecular gravity decreases with temperature.

Phase Transition

There are two pressure/attractive forces: the molecular gravity and the normal pressure. Since
both components are attractive, molecules would be squeezed together. There is no repulsion
force for phase transition to gas direction, which is clearly incorrect. So, we postulate that the
orbital repulsion force plays this role. As shown in the helium model, the orbital repulsion
increases with molecule size, which in turn increases with temperature. With the understanding
of both trends, we can semi-quantitatively analyze the phase transition in terms of temperature.

The state of matter is determined by the interaction of three forces: the orbital repulsion, the
molecular gravity and pressure. The three forces are shown in Figure 6B. Orbital repulsion
increases with temperature as shown with the orange curve. The pressure is shown as the
horizontal line in light blue. The molecular gravity is shown in a blue curve near the bottom of
Figure 6B. The net force between molecules are platted in a black curve. So, the net force is
attractive at low temperatures, which becomes repulsive gradually with increasing temperature.
The net force determines the phase transition from solid to fluid and supercritical fluid.
Figure 6, Phase transition and intermolecular force.

Orbital repulsion force also plays an important role in phase transition. In Figure 6B, the
solidifying line is the transition from solid to fluid where the orbital repulsion force ​Fs breaks
even with the molecular gravity/attraction force ​Fg​. To the left of this point, molecular attraction
is greater than orbital repulsion. The attraction is able to hold molecules together against the
repulsion in a fixed frame. Solid is observed. To the right, orbital repulsion overcomes molecular
attraction. Molecules are pushed apart and broken into fluid. The temperature at this point
becomes the transition point between solid and fluid phases. At a higher pressure, it requires a
higher repulsion force to break the solid. So, the solidifying line curves to high temperatures with
increasing pressures because orbital repulsion increases with temperature.

Note, the net force may still be attractive to the right of the solidifying point. Even though the
force is not enough to hold against the orbital repulsion, but still provides the attraction between
molecules to create viscosity in fluid and surface tension on the liquid. The attraction is also
responsible for the phenomenon of capillarity, where liquid seems defying gravity. Orbital
repulsion occurs only when there is orbital jump. On the surface of liquid or solid, orbital
repulsion kicks and ejects molecules into the air in the process known as vaporization or
sublimation. Molecular attraction decreases with increasing temperature. At a temperature of
supercritical point, Figure 6A, the attraction ​Fp breaks even with repulsion ​Fc​. Beyond this
point, repulsion force is greater. There is no attraction force to provide viscosity between
molecules in fluid and no surface tension on liquid. Because of that, there is no latent heat in
phase transition between liquid and gas, therefore the transition is not distinctive, a state also
known as supercritical fluid.​[17-19]

In a sense, pressure provides additional force to restraint molecules together. Liquid is just
pressured gas. In fluid, as pressure increases, molecules are forced to stay at a closer range.
As pressure increases to a point, the liquefying pressure, Figure 6A, intermolecular attraction is
able to glue the molecules together into liquid. The liquefying line curves up with temperature
because molecular repulsion increases with temperature. The liquefying curve does not extend
beyond the critical point because there is no distinction between liquid and gas.

Conclusions

Phase transition is determined primarily by orbital repulsion and molecular gravity/attraction.


Molecular attraction is responsible for glue molecules together in liquid and solid. The orbital
repulsion works against the attraction to break solid into fluid as temperature increases. The
transition from solid to fluid is the point where orbital repulsion force break even with the
molecular attraction force. The residual net attractive force beyond the solidifying curve provides
the force for viscosity in fluid and surface tension on the liquid. Beyond the critical point, there is
no attraction force to provide viscosity in fluid. Since there is no attraction at all, there is no
latent heat and no distinction between liquid and gas, known as supercritical fluid. Liquid is just
a special gas under pressure.

References

1. Margenau, H.; et al. (1969). “​Theory of inter-molecular forces”​, International


Series of Monographs in Natural Philosophy, Pergamon Press.
2. Oxtoby, D.W.; et al. (2012). “​Principles of modern chemistry” (7th ed.). Belmont,
Calif.: Brooks/Cole Cengage Learning.
3. Bader, R.F.W.; et al. (1965). "​The Ionic Bond​". Journal of the American Chemical
Society. ​87​ (14): 3063–3068. doi:​10.1021/ja01092a008​.
4. Dzyaloshinskii, I.E; et al. (1961). "​General theory of van der waals' forces​". Soviet
Physics Uspekhi. ​4​ (2): 153.
5. Zheng, Y.; et al. (2011). "​Lifshitz Theory of van der Waals Pressure in Dissipative
Media​". Phys. Rev​. ​A. ​83​ (4): 042504. arXiv:​1011.5433​.
6. Tadmor R. (2001). "​The London-van der Waals interaction energy between
objects of various geometries"​ . Journal of Physics: Condensed Matter. ​13 (9):
L195–L202.​ ​doi:​10.1088/0953-8984/13/9/101​.
7. Roller, D.; Roller, D.H.D. (1954). “​The development of the concept of electric
charge: Electricity from the Greeks to Coulomb”​. Cambridge, MA: Harvard
University Press. p. 79.
8. Stewart, J. (2001). “​Intermediate Electromagnetic Theory”.​ World Scientific. p. 50.
9. Baigrie, B. (2007). “​Electricity and Magnetism: A Historical Perspective”.​
Greenwood Press. pp. 7–8.
10. Comins, N.F.; Kaufmann, W.J. (2008). “​Discovering the Universe: From the Stars
to the Planets”​ . MacMillan. p. 347.
11. Weinberg, S. (1972). “​Gravitation and cosmology”​. John Wiley & Sons. p. 194.
12. Weisstein, E.W. (2007).​ ​"F ​ ifth Force​"​. World of Science. Wolfram Research.
13. Schrödinger, E. (1926). ​"​An Undulatory Theory of the Mechanics of Atoms and
Molecules​"​. Physical Review. ​28​ (6): 1049–1070.​ ​doi:​10.1103/PhysRev.28.1049​.
14. Liu, J.Z. (2019), “​Uncovering the Mystery of Superconductivity​”, Stanford
University.
15. Bondi, A. (1964). "​Van der Waals Volumes and Radii​". ​J. Phys. Chem​. ​68 (3):
441–451. doi:​10.1021/j100785a001​.
16. Rowland, R.S.; et al. (1996). ​"​Intermolecular Nonbonded Contact Distances in
Organic Crystal Structures: Comparison with Distances Expected from van der
Waals Radii​"​ ​J. Phys. Chem​.​ ​100​ (18): 7384–91. doi:​10.1021/jp953141​.
17. Székely K. (2014). ​"W ​ hat is a supercritical fluid?"​ ​. Budapest University of
Technology and Economics.
18. Padrela, L.; et al. (2009). "​Formation of indomethacin–saccharin cocrystals using
supercritical fluid technology​". European Journal of Pharmaceutical Sciences. ​38
(1): 9–17. doi:​10.1016/j.ejps.2009.05.010​.
19. Ye, X.R.; et al. (2003). "​Supercritical fluid fabrication of metal nanowires and
nanorods templated by multiwalled carbon nanotubes​". Advanced Materials. ​15
(4): 316–319. doi:​10.1002/adma.200390077​.

You might also like