1122574 (1)
1122574 (1)
[Document subtitle]
Abstract
[Draw your reader in with an engaging abstract. It is typically a short summary of the document.
When you’re ready to add your content, just click here and start typing.]
Farah
[Email address]
Abstract
A detailed overview of the origins and possible derivations of the ideal
gas law are given in this lecture text. The tacit assumptions in the usual
textbook derivations are identified and ways of thought are proposed to
justify their use. It is shown that the solution of the Schrödinger equation for
the particle in a box model is very closely connected to the ideal gas law. A
detailed derivation is presented on how equipartition and the Boltzmann
distribution lead to the velocity and speed distributions observable in ideal
gases. Major emphasis is given to the mathematical consequences of the lack
of direction dependence in these properties. General formulas are introduced
and proved to describe the thermodynamics of ideal gases. During the
derivations, a method for students to employ computer-aided symbolic
calculations using the Wolfram Alpha web-based service is also described.
Introduction
The concept of ideal gas and derivation of the ideal gas law
Some confusion may arise from the fact that the quantities' velocity and speed
are not always distinguished clearly. In this text, we will strictly stick to the
convention that speed
(s) is the size of the velocity vector. Therefore, it is a scalar quantity without
direction, and it cannot be negative. Speed can be calculated from the velocity
vector using a simple geometric argument (Pythagorean theorem):
S=√ v2 + v2 + v2.
x y z
Note that it is impossible to obtain the velocity from the speed alone, and
confusion often arises from the fact that the symbol v is used for speed as well.
The law of momentum conservation is straightforward in physics and can be
applied to non-elastic collisions as well. The law of energy conservation is
somewhat different. There are many different kinds of energy (e.g., electric,
rotational, vibrational), but for an ideal gas, the only thing we care about is
kinetic energy. This can be calculated by multiplying the mass with the square
of the speed and then dividing by two, or also with the components of velocity:
Therefore, kinetic energy is a scalar quantity (this is of course, valid for all kinds
of energy). The assumption behind an elastic collision is that the particles do not
change any of the other kinds of energy in it: they do not get irreversibly
deformed, and if they rotate or vibrate, they do so in the same way after the
collision. This means that the sum of the kinetic energies of the two colliding
particles is the same before and after the encounter.
Two particles that are involved in a collision have six velocity components
combined. Even if we know all of them before the collision, they will take new
values after the event, so six unknowns arise. There will be three equations from
the conservation of momentum (one for each component) and one additional
equation from the conservation of energy. This means four equations altogether,
which is not enough to define six unknowns. It is prob- ably good to remember
that considerations beyond the conservation laws are needed to calculate the
outcome of an elastic collision, but fortunately, these are not necessary in
dealing with the properties of ideal gases.
Many statements can be made about an ideal gas [5–8, 16–18], but the following
is already sufficient as a physical definition: the only possible interaction
between the particles is the elastic collision. Note that the concept is a scientific
abstraction, and a substance in itself cannot be called an ideal gas. For example,
hydrogen and helium are not ideal gases in a strict sense. The precise statement
is always that a certain substance behaves like an ideal gas under certain
conditions. For hydrogen and helium, certain conditions include room
temperature and atmospheric pressure. Even gold behaves like an ideal gas at
3500 K and 1 Pa.
Notably, the definition given above does not require identical particles in the
gas. It can be a mixture of different substances. Most of the intensive physical
properties of an ideal gas depend on the chemical composition. An obvious such
property is molar mass (M); additional examples include density (ρ) and molar
heat capacities (CV,m and Cp,m). There are few exemptions: for example, pressure,
temperature, molar volume and concentration. These are the ones used in Eqs.
2 and 3, and this is a pivotal point in ideal gas behavior: one does not need to
know the identity of the gas to work with these selected properties. This is the
very reason why the forms given in Eqs. 2 and 3 are highly preferred to all other
possibilities. Also, this was the essential discovery in Avogadro’s law [19, 20]. It
should be noted that the elastic nature of the collisions between particles is
emphasized in most textbooks (p. 11 in [8]). However, from a strict physical
point of view, this is only true if the particles are structureless (e.g., noble gas
atoms). In multiatomic molecules, there are additional internal forms of
molecular motion that can store energy, most importantly rotations and
vibrations.
In a collision, it may happen that these are increased at the expense of the overall
kinetic energy, or their decrease provides extra kinetic energy. Fluctuations of
this kind are possible, but the equipartition theorem guarantees that the two
conservation laws behind elasticity will be valid statistically over a large number
of collisions. Also, an individual collision with the wall seems to violate the
conservation of momentum as the particle bounces back (the direction of the
velocity changes) without any counterpart, as the wall is stationary. This picture
is still useful as the wall is considered to have such a large mass that gaining a
little bit of momentum from a molecular collision will not change its state of
motion. In addition, similarly to the intermolecular case, the overall efect of a
large number of molecular collisions with all parts of the container wall will give
an average of zero change in the overall momentum.
A simplistic view of pressure
This could be difficult to distinguish from the product of two symbols. Similar
notation customs are used for differentials with the letter d and for partial
differentials with the symbol ∂.) Some knowledge of the velocities of the
particles is necessary to do this, but only in the direction that is perpendicular
to the chosen surface A. Let the symbol for this direction be x. Consider particle
M1, whose velocity component in the direction x is vx. If vx is the same as the
average of this property for all particles, then half of the articles in the
rectangular box of base A and depth vxΔt collide with A in this interval. The
reason why only half of the particles do this is that in truly random motion, the
other half moves away from A and will not collide within time Δt. The volume of
the box is AvxΔt, so using the concentration (c) of the gas, the amount of
substance for the colliding particles is cAvxΔt/2. The number of such particles is
obtained by multiplication with the Avogadro constant NA (= R/k) to yield
NAcAvxΔt/2.
Our main interest is the pressure, which is defined in physics as force divided by
the area upon which it acts. Furthermore, force is described in Newton’s laws as
the overall momentum change divided by the time interval in which it occurs.
When particle M1 collides with the wall, it is a very special kind of collision: a
bounce (meaning that the mass of the wall is so much higher than the mass of
the molecule that the wall does not move at all). The physical laws of bouncing
require that the velocity of the particle in direction x changes from vx to − vx,
while other components remain unchanged. This is shown by a dotted black
arrow in Fig. 1. So, the overall change of velocity in direction x is |− vx− vx |= 2vx
(the | | signs represent absolute value). If the mass of the particle is m, this
means a momentum change of 2mvx. Therefore, the pressure is obtained by
multiplying this momentum change (2mvx) with the number of particles
colliding (NAcAvxΔt/2) and then dividing by the time interval (Δt) and the
surface area (A):
It can be seen that the surface area and the time interval conveniently disappear
during simplifcation, which is readily rationalized by the physical observation
that the pressure of a sample of gas is independent of where and when it is
measured. As already mentioned, the simple sequence of thought leading to
Eq. 6 can be found in many textbooks as a derivation of the ideal gas law,
although the work used as a reference here [8] was notably improved compared
to its original version with the publication of new editions. However, this
derivation is seldom taken to completion. Instead, Eq. 6 is typically simply
compared to the ideal gas law (whose validity is assumed before the end of the
derivation!), and a formula is obtained for vx. This typical approach is most
unfortunate as there is a simple line of thought that can be used to complete the
derivation easily. This relies on the equipartition principle: first, the term mvx 2
/2 is recognized as the contribution of the motion in direction x (Ex) to the
average kinetic energy of the particles; then, the equipartition principle is
applied to replace this term with kT/2:
It is seen that we arrived at the form of the ideal gas law written in Eq. 3, so the
derivation is complete now. Historically, this equation could be used to define
the universal constants k and R. As already pointed out, textbooks typically do
not bring the derivation to completion. In the absence of the use of equipartition,
this is not even a proper derivation of the ideal gas law. However, there are other
problems with the presented sequence of thought as well. The reader is asked to
stop for a moment at this point and try to find lapses of logic in the sequence of
thought presented above in Sect. "A simplistic view of pressure." Normally, few
people are inclined to do this as textbook knowledge is assumed to be beyond
any doubt. Unfortunately, this is far from the truth. Similar to all other
documents, textbooks also contain errors, which may range from minor typos
or regrettable omissions to outright fallacies. The next subsection will list a (not
necessarily complete) collection of such lapses of logic in this sequence of
thought. None of them are entirely beyond repair, but they necessitate
additional remarks at the very least. If the reader was successful in identifying
one or more of them, it is a favorable sign of the ability to think scientifically.
Direction
Consider particle M3 in Fig. 1. It is not within the rectangular box for which we
have made the count, but it does collide with A within the given time interval.
Or consider particle M2. It is within the box; it moves toward A but collides with
the wall outside A. So, M3 is not counted but contributes to the pressure,
whereas M2 is counted but does not, in fact, contribute. Refuge from this
problem can be found in the fact that there is a very large number of molecules
in the system, and their motion is completely random, which gives a good reason
to hope that the effects of entering and leaving particles will cancel each other
out exactly.
Collisions
Consider particles M5 and M6 in Fig. 1. They are both within the rectangular
box of interest, and they are both moving toward A. However, they collide with
each other before hitting the wall (only the left view of Fig. 1 shows this fact
clearly) and change the direction of their motion. There are several good
remedies for this problem. The first is noticing that the conservation of
momentum law is valid separately for the different components of the vector. As
the collision cannot change the sum of the momenta of the two particles in
direction x, this momentum will reach surface A and exert a force despite the
intermolecular collision. The second is based on the fact that all appearances of
the time interval Δt are cancelled by the operations in Eq. 6. That means we are
free to select a Δt that is short enough to avoid any such collisions. The third way
out is a highly counterintuitive one: we did not clearly specify this earlier, but
we imagine the particles to be mass points at the first approach [21]. Roughly
speaking, points do not have measurable sizes in any direction, so they do not
collide. To satisfy even the most pugnacious mathematician, the probability of
two mass points colliding in this model is zero in any finite time interval. A
simultaneously amusing and amazing further fact in probability theory is that
an event with zero probability is still not impossible [22]. This line will not be
pursued here any further: understanding the quantitative description and the
consequences of collisions between particles in an ideal gas needs such scientific
depth that a separate article will be devoted to it.
Relativity
The subtitle is honest here but may still be misleading. No, we will not argue that
Einstein’s relativity theory should be brought into the discussion in any way.
What we simply mean is that reference points are needed to measure velocity or
speed. These very often take the form of a coordinate system. The derivation
presented to calculate the pressure tacitly assumed that the wall itself does not
move; only the particles change their positions. So, the coordinate system is
fixed to the container of the gas sample. Still, we were notably free to choose a
direction to be designated x. There will be some consequences of this freedom,
which we will discover in a later section. Also, at this point, it may be useful to
state that most of the derivations presented in this article are only valid if the
center of mass of the gas sample does not move relative to the container. In other
words, transferring the results to the case of following gases would need
additional mental work.
Averages
Now, compare the velocities of particles M1, M2, M3, M4, M5, and M6 in Fig. 1.
Of course, they are different because the directions of movement are different.
However, a closer look will reveal that even the speeds are different. This is a
natural state of affairs in a gas sample: the hypothetical state in which all the
particles move with the same speed is definitely outside the boundaries of
classical thermodynamics.
The speeds and velocities of different particles vary then; yet, we used a single
value for vx. First, we tried to define it as a bona fide average and then attempted
to update the definition to require an average for those particles that move
toward A but warned that this would create even more problems. Now, it is time
to deliver on this promise. The problem is averaging itself, which can be
demonstrated in an example that requires some numerical calculations.
Consider two gas particles moving at different speeds, for example, two helium
atoms, one moving at 100 m s −1, the other one at 500 m s −1. It is a very natural
thing to state that the average speed of the two molecules is 300 m s−1 in this
case. Now, calculate the kinetic energy of the two particles (E=ms2 /2). The
mass of a helium atom is m = M /NA =0.00400 kg mol−1/6.02× 1023
mol−1=6.64× 10−27 k g, so the kinetic energy of the one traveling at 100 m s−1
is 3.32× 10−23 J=33.2 yJ (y stands for yocto, a rarely used prefx for 10−24) or
20.0 J mol −1. The kinetic energy of the helium atom moving at 500 m s −1 is
830 yJ (or 500 J mol−1), so the average energy of the two atoms is 432 yJ or
260 J mol−1 (note that the decimal part is not given, as this would run against
the established scientifc use of signifcant digits). If one calculates the kinetic
energy for the average speed of 300 m s −1, the result is 299 yJ (180 J mol−1).
There is no mistake here: the average energy cannot be calculated from the
average speed! Or at least not from this average. When discussing the
description of ideal gases, many textbooks introduce the concept of root mean
square speed, which is, as the name implies, the square root of the average of
the squares of the two speeds, srms=((1002+ 5002 )/2)1/2= 361 m s −1. If the
average energy is calculated from srms, the correct value of 432 yJ (or
260 J mol−1) is achieved.
The origin of the problem is quite deep. When we consider a large collection of
things (such as molecules) whose properties differ, there is no natural way of
calculating ‘the average.’ There are several different averages [22]. The one that
seems the most natural to many users is called the arithmetic mean: we add all
the properties and then divide by the number of things considered. The root
mean square has the mathematical name of quadratic mean, and there are
others as well. When we try to characterize the properties of a bunch of particles
with different individual properties, it is difficult to know which of the means is
needed in a formula. It is often necessary to explore the probabilities at which
certain values occur in a physical property. This idea will be elaborated in later
sections.
A careful reader might have already spotted an apparent contradiction. The end
of Sect. "The notion of an ideal gas" defined the ideal gas as a sample for which
the only possible interaction between the particles is the elastic collision. On the
other hand, Sect. "Collisions" said that if the particles of an ideal gas are thought
to be mass points, they collide with zero probability. This may simply be a
language issue, but by formal logic, the absence of collisions actually still
satisfies the ‘only possible interaction…’ criterion. The easiest way to exclude the
possibility of collisions is to have only one particle in a container.
Could one particle make sense as an ideal gas? It certainly collides with the walls
of the container, so it must exert some pressure. It is more difficult to imagine
how we could attribute temperature to a single particle. Nevertheless, it is worth
putting some effort into this as it is an exceptionally well-characterized system
in quantum mechanics for which a reasonably detailed mathematical
description is given in textbooks as well. It is called particle in a box (p. 264 in
[8]).
All the physical properties of a particle in a box in thermodynamic equilibrium
can be obtained from its stationary wave function Ψstat, which is the solution of
the stationary Schrödinger equation:
In the particle in a box model, these additional equations are called boundary
conditions, and they reflect the physical fact that the particle cannot occur in or
outside the walls of the container. For a rectangular box of dimensions a×b×c,
these boundary conditions can be stated in a simple manner:
As already stated, finding the solution shown in Eqs. 10–11 is quite a demanding
task, but once known, it is easy to make sure that it satisfies the equations by
simple substitutions.
As a next step, we can calculate the force acting on one of the walls of the box.
Since force does not appear in any of Eqs. 8–11, it must be connected to the
existing physical properties somehow. For this, it is useful to recall a piece of
basic physical knowledge: energy is the ability to do work, and work can be
calculated as the product of the force and the displacement in the direction of
the force. So, when the wall of the rectangular box in direction x is moved a bit,
the force can be calculated as the ratio of the energy change and the
displacement in direction x. In more formal mathematical language, it is said
that the force is the derivative of the energy regarding a change in the side length
a. As the energy is already known as a function of a, this calculation can be
brought to an easy conclusion:
In an ideal gas, the particles are independent of each other, and if they are
considered mass points, they do not collide with each other. So, for N particles
(N=nNA), the overall force is obtained by simply summing the contributions
from the individual particles:
At this point, we can use the equipartition theorem again. The fact that the
average energy per degree of freedom for a single molecule is kT/2 can be put
into a mathematical equation for our case as follows:
The pressure is the absolute value of force divided by the area of the
wall perpendicular to direction x, which is easily given by the product
of the two sides (bc):
So, the most common form of the ideal gas law (Eq. 1) can be derived
directly from the quantum mechanical model of a particle in a box.
Thus far, this derivation was presented for a rectangular box. However,
the ideal gas law, in fact, has no limitations on the shape of the box in
which the gas is stored. The previous considerations can be somewhat
extended into containers of different shapes, but these rely on solving
the stationary Schrödinger equation for different boundary conditions.
While a particle in a (rectangular!) box is a typical textbook model in
quantum mechanics, other cases are seldom handled. This article gives
two more examples: the case of a cylindrical box and the case of a
spherical box.
A cylindrical box is imagined in a way that the cross section of the box
in the yz plane is a circle of radius R, whereas in the two other
perpendicular planes, it is a rectangle. The height of the cylinder is
denoted a; its meaning is exactly the same as that of c for a rectangular
box. The details and the solution of the Schrödinger equation are given
in the Supplementary Information. What is important is the formula
obtained for the possible quantized energy levels of a particle. This
contains an nx quantum number analogous to that seen in Eq. 9. There
are two additional quantum numbers that arise for motion in the
direction perpendicular to the x-axis, but they do not appear in the
energy formula. Instead, a series of infinitely many isolated values
denoted Zml appear, which represent the independent variables at
which the value of the Bessel function of the first kind is 0. This might
sound like a frightening definition, but for our purposes, it is a constant
whose values will not even be needed. The full energy formula is:
The same can be done in the direction z, which would give the pressure
acting on the bases of the cylinder. The force here is calculated by
derivation with respect to a:
The case of the spherical box (with radius R) also brings some
difficulties in solving the stationary Schrödinger equation; the solution
is shown in the Supplementary Information. The quantized energy
levels are given by a formula that includes a constant Zl, which
represents the independent variables at which the value of the
spherical Bessel function of the first kind (jl ) is 0. Again, the actual
values of Zl will not be needed:
The force acting upon the spherical wall is obtained from a derivation
similar to the previous ones:
Now, R characterizes the boundaries of a three-dimensional
movement, so the average energy should be taken as 3×½kT. For
calculating the pressure, the surface of the sphere is needed now:
In all cases (Eqs. 15, 18, 20, and 23), the validity of the ideal gas law is
confirmed for containers of different shapes. However, it must be
recognized that only simple shapes can be handled by this method, and
even then, the derivation is notably burdensome as it involves finding
the solutions of the stationary Schrödinger equation with different
boundary conditions. So, overall, this line of thought shows quite
nicely the fundamental origins of the ideal gas law but is very
impractical as a proof and becomes close to impossible for more
complicated container shapes. So, in the next section, a line of thought
will be sought where the container shape does not matter so much.
There are two unfortunate features of Eq. 26. First, the use of the letter
F to denote a cumulative probability function is almost exclusive in the
literature. In this article, F was already used for the physical property
of force, which is also an almost exclusive notation in science. The only
remedy that can be offered is that F used for a cumulative probability
function is always followed by parentheses, and this is never done for
force. The second mishap connected to Eq. 26 is mathematical: the
infinite sums appearing in it are quite difficult to calculate. It is
absolutely necessary to take an approximation step, which is based on
the observation that at temperatures > 10 K, the P(nx) probabilities
corresponding to two neighboring nx values are barely different. The
consequence is that the possible nx values can be approximated well
with a continuum instead of quantized, discrete values. The continuum
equivalent of addition is integration, so we can replace the infinite
sums in Eq. 26 by infinite integrals, which is more helpful than it
seems at first sight because there is a huge, time-honored experience
in calculating infinite integrals in mathematics. This transition uses
the fact that the histograms that would be drawn with an interval of 1
(Δi = 1) to calculate the infinite sums can be approximated well by the
area under the continuous curve drawn to the top of the intervals:
In this case, a very good approximation can be given for the cumulative
probability function:
kjhkjhj
Fig. 2 Scaled cumulative distribution Fig. 3 Scaled probability density
function for component x of the velocity function for component x of the
of a particle in an ideal gas velocity of a particle in an ideal gas
It is likely that the curve shown in Fig. 3 looks very familiar to most
readers with some scientific background. The curve itself is called a
Gaussian curve; the distribution it represents is called a normal (or
Gaussian) distribution. The general form of its probability density
function is this:
The two parameters in the general form are called expectation (μ) and
standard deviation (σ). A comparison of Eqs. 32 and 33 show that
these parameters have the following value for the distribution of
component x of the velocity of a particle in an ideal gas:
Using these rules in Eq. 35 reveals that the resulting vθ,φ random
variable, constructed by the linear combination of vx, vy and vz, is
normally distributed with an expectation of zero. Its standard
deviation σθ,φ can also be calculated from the standard deviations of vx,
vy and vz (σ = k½ T½ m−½):
The last part of Eq. 37 is the classical formula for impingement rate,
indeed (p. 711 in [8], but named ‘collision fux’ there).
There are two integrations in this formula. The inner one (ending in
dvx) calculates the number of species that show vx≥x/Δt at a fxed
distance of x from the wall (the ones which will collide within a time
interval of Δt), whereas the outer integration (ending in dx) sums up
the contributions from all particles at any distance x from the wall.
This derivation gave the ideal gas law in a form that was
already shown in Eq. 3.
The shapes of these two energy distributions are shown in Figs. S1 and
S2 in the Supplementary Information. The arithmetic mean of the
energy can be calculated in a way that is fully analogous with the one
shown for Eq. 39:
This formula verifies that the velocity component and energy
distributions obtained here are in agreement with the equipartition
theorem. Furthermore, a comparison of Eqs. 41 and 45 uncovers the
following relation, which may be unsurprising but still notable:
The arithmetic mean for the square of the relative velocity components
turns out to be twice the value calculated for the analogous quantity
measured in the other coordinate system:
Speed distribution
Now, we can turn our attention to the speed of the particles. The
distributions of the velocity components are already known; so is
Eq. 4, which connects them to speed. From a probability theory view,
what needs to be calculated is the square root of the sum of the squares
of three identical, independent normal distributions. Fortunately, this
distribution is known on its own right; it is called chi distribution with
three degrees of freedom. The easiest way to discover this is simply to
read a good textbook on probability theory. Going into details about
the derivation and properties of this distribution would lead too far
into the kingdom of mathematics, but one sequence of thought will be
shown here. To fnd a formula for the speed distribution, we should
remember that such changes on independent variables are best
attempted using the cumulative distribution function. To obtain the
probability that the speed of a particle is smaller than s, we need to add
the product of probability densities f(vx), f(vy) and f(vz) for all velocity
vectors for which vx 2+vy 2+vz 2
One of the central issues in the kinetic molecular theory for ideal gases
was the (average) kinetic energy of gas particles. Therefore, it is natural
to seek connections with the classical thermodynamic functions such
as internal energy, enthalpy and entropy [23, 24]. As the frst law of
thermodynamics makes a statement about internal energy, it is the
logical beginning of the discussion. We will keep our earlier stated
preference for intensive physical properties, so molar internal energy
(Um) will be used. The temperature dependence of this quantity at
constant molar volume is described by an important derivative called
molar heat capacity (CV,m):
Then, one could obtain CV,m as the ratio of CV and n. The assumption
of a closed system is not always necessary for the common equations
used in thermodynamics; some of them are valid for open systems as
well. The otherwise necessary care about the amounts of substances in
relation to other extensive quantities only appears in textbooks when
the discussion of chemical potentials begins. To avoid these problems,
this text will keep sticking to the use of intensive quantities
consistently.
Unfortunately, CV,m is not specifed by the fact that a gas follows ideal
behavior. It is not only that a value is not provided for CV,m, but it is
not even necessarily independent of temperature! This is an important
point: many formulas found in textbooks for the thermodynamics of
ideal gases tacitly assume a temperature-independent CV,m, which
may sometimes be a source of confusion. In this text, we will deal frst
with some cases where this assumption is not necessary. It is usually
advantageous to consider molar internal energy as a function of
temperature and molar volume. This is simply a matter of selection:
Gibbs’ phase rule ensures that giving two of the properties of a one-
component, onephase system defnes all others. In this case, the most
common choice is out of pure convenience: formulas tend to be
reasonably simple with this selection.
References
1. James SW (1928) The discovery of the gas laws. I. Boyle’s law. Sci.
Prog. 20th Cent. 23:263–272 https://www.jstor.org/stable/
43428666. Accessed 19 Nov 2024
2. James SW (1929) The discovery of the gas laws. II. Gay-Lussac’s law.
Sci. Prog. 20th Cent. 24:57–71 https://www.jstor.org/stable/
43428731. Accessed 19 Nov 2024
3. James SW (1930) The discovery of the gas laws. III. The theory of
gases and van der Waals’ equation. Sci. Prog. 20th Cent. 25:232–239
https://www.jstor.org/stable/43430284. Accessed 19 Nov 2024
4. Scerri ER (2006) The periodic table: its story and its signifcance.
Oxford University Press, Oxford. https://doi.org/10.1093/oso/
9780190914363.001.0001 5. Davis WM, Dykstra CE (2012) Physical
chemistry. A modern introduction, 2nd edn. CRC Press, Boca Raton.
https://doi.org/ 10.1201/b11675
17. Laugier A, Garai J (2007) Derivation of the ideal gas law. J Chem
Educ 84:1832–1833. https://doi.org/10.1021/ed084p1832
18. Woody AI (2013) How is the ideal gas law explanatory? Sci Educ
22:1563–1580. https://doi.org/10.1007/ s11191-011-9424-6
20. Held L (2017) Avogadro’s Hypothesis after 200 Years. Univ J Educ
Res 5:1718–1722. https://doi.org/10.13189/ujer.2017. 051007
23. Levine S (1985) Derivation of the ideal gas law. J Chem Educ
62:399. https://doi.org/10.1021/ed062p399.1