You are on page 1of 22

Vol.

5, 2020-14

A Barometric Formula without the Hydrostatic Pressure Assumption

Hugo Hernandez
ForsChem Research, 050030 Medellin, Colombia
hugo.hernandez@forschem.org

doi: 10.13140/RG.2.2.20093.49126

Abstract

Barometric formulas are important mathematical equations used to understand and predict
the behavior of the atmosphere pressure at different altitudes. Since the first development by
Pierre-Simon de Laplace in the 18th century, the fundamental assumptions leading to
barometric formulas have been considering air as an ideal gas at steady-state, and considering
atmospheric pressure as a hydrostatic pressure following Pascal’s law. Being rigorous however,
gases do not follow Pascal’s law since the molecules are on average so far from each other that
they cannot transmit the weight of their neighboring molecules in the vertical direction. For
this reason, a new barometric formula has been derived without recurring to the hydrostatic
pressure assumption. Instead of Pascal’s law, the conservation of momentum is used to
describe the effect of gravity on the vertical molecular density profile. Then, after determining
the temperature profile (which can be derived by solving the energy conservation equation, or
can be empirically obtained), the molecular density profile can be solved, and the vertical
pressure profile can be directly obtained from the ideal gas equation. The barometric formula
obtained, which is almost equivalent to the current barometric formula used by the standard
atmospheric model (the US Standard Atmosphere of 1976), was tested considering a set of
experimental barometric measurements reported from different locations worldwide. Even
though only a slight difference is obtained, the new expression no longer requires assuming
atmospheric pressure as hydrostatic. The wide success of previous barometric formulas can be
explained by the fact that the pressure drop predicted by the conservation of momentum
deviates by less than from Pascal’s law. Finally, a multicomponent model of air was
considered, which allows the estimation of atmospheric composition changes with altitude.

Keywords

Atmosphere, Barometric formula, Energy conservation, Hydrostatic pressure, Lapse rate,


Molecular distributions, Momentum conservation, Pascal’s law, Troposphere

26/08/2020 ForsChem Research Reports Vol. 5, 2020-14 (1 / 22)


www.forschem.org
A Barometric Formula without the
Hydrostatic Pressure Assumption
Hugo Hernandez
ForsChem Research
hugo.hernandez@forschem.org

1. Introduction

In a recent publication, Lente and Ősz [1] explained in great detail different lines of thinking
resulting in the most common barometric formulas used to describe changes in atmospheric
pressure with altitude. Interestingly, all the approaches considered provide fairly good
approximations for describing the pressure profile, even though some of them were based on
different principles.

They classified the models according to the temperature profile at different altitudes into:
Constant temperature models and variable temperature models. While the constant
temperature assumption of the whole atmosphere is clearly unrealistic, it is the starting point
of the derivation of the formulas. Once the temperature profile is known, the extension as a
variable temperature model is relatively straightforward. Among the constant temperature
models considered, only the model based on the statistical thermodynamics approach cannot
be extended as a variable temperature model.

Excluding the statistical thermodynamics approach, all other approaches presented for
deriving the barometric formula rely on the hydrostatic pressure assumption, or equivalently,
on the validity of Pascal’s law for atmospheric air. In other words, these approaches assume
the validity of the following expression:

̃
(1.1)

where is the atmospheric pressure at altitude , ̃ is the air molecular density at altitude ,
is the average molecular mass of air, and is the gravitational acceleration (also at
altitude ). The negative sign indicates that the atmospheric pressure decreases with altitude.
Eq. (1.1) basically expresses that the atmospheric pressure is caused by the weight of the mass
of air above a surface.

Pascal’s law was originally proposed for describing pressure differences in liquids [2], and has
been shown unsuitable for gases [3,4]. In gases, the increase in pressure as the altitude
decreases is not directly caused by the weight of the column of gas, as indicated by Pascal’s
law. In reality, the additional pressure is caused by an increase in mass density and in vertical
molecular velocities as altitude is reduced. Such effect is the result of the gravitational
acceleration, but not of weight transmission along the gas body in the vertical direction. Even
more, the increase in atmospheric pressure close to the surface is even larger than the pressure
increase expected by the increased weight of air. However, under certain conditions gases can
behave approximately as a Pascal fluid [3].

26/08/2020 ForsChem Research Reports Vol. 5, 2020-14 (2 / 22)


www.forschem.org
A Barometric Formula without the
Hydrostatic Pressure Assumption
Hugo Hernandez
ForsChem Research
hugo.hernandez@forschem.org

Thus, even though common barometric formulas are good approximations for describing the
vertical atmospheric pressure profile, they are based on a wrong assumption: The validity of
the hydrostatic assumption in gases.

The purpose of this report is presenting an alternative derivation of a barometric formula


without assuming hydrostatic pressure, but using conservation equations instead.

2. Molecular Description

The common derivations of the barometric law assume atmospheric air to be an ideal gas. For
typical environmental conditions, and for practical purposes, this assumption is reasonable. An
ideal gas is a low-density molecular system where each molecule, on average, experiences a
net intermolecular force of zero. However, in the case of atmospheric air the average net force
is not zero because of the gravitational force acting on all molecules. In addition,
intermolecular forces acting between the ideal gas molecules and the solid surface play an
important role in the macroscopic mechanical stabilization of the atmosphere.

On the other hand, the fact that the average intermolecular force (between ideal gas
molecules) is zero does not imply that molecules do not (or cannot) collide. On the contrary,
collisions between ideal gas molecules, while relatively infrequent compared to other systems,
are very important for chemical kinetics and transport phenomena [5].

A schematic representation of an arbitrary fraction of the atmosphere is presented in Figure 1.


Air molecules move randomly in the horizontal direction, with an average zero velocity. This
also implies that there is a homogeneous molecular distribution in the horizontal direction.
However, in the vertical direction, the velocities are greatly influenced by gravity, which result
in a steady vertical distribution profile. Of course, no molecular system is static. However, an
average steady-state can be obtained, which will be considered in the development of the
model.

The steady-state assumption implies that the molecular concentration profile remains
constant, despite the fact that molecules are in permanent motion. Under this assumption, the
net flux of molecules crossing an imaginary horizontal boundary is zero, but the individual
upward and downward fluxes are never zero. This is illustrated in Figure 2.

26/08/2020 ForsChem Research Reports Vol. 5, 2020-14 (3 / 22)


www.forschem.org
A Barometric Formula without the
Hydrostatic Pressure Assumption
Hugo Hernandez
ForsChem Research
hugo.hernandez@forschem.org

Figure 1. Molecular representation of atmospheric air. Black dots represent air molecules. A
molecular concentration profile develops vertically as a result of gravity.

Figure 2. Molecules crossing an imaginary horizontal boundary located at height in a short


period of time. Green arrows represent the vertical velocity of molecules crossing the boundary
in the upward direction. Blue arrows represent the vertical velocity of molecules crossing the
boundary in the downward direction.

The distribution of vertical molecular velocities at each particular height will be influenced by
gravity but also by molecular collisions. Considering the large number of collisions taking place
over short periods of time, and following the ideas of Maxwell [6] and Boltzmann [7], a
reasonable assumption is that the vertical molecular velocities at each height are normally
distributed, with a mean value of zero. At the surface, by assuming perfectly elastic collisions,
the steady-state requirement is still satisfied.

Considering air as a multicomponent system, the vertical molecular speed (the magnitude of
the molecular velocity) of species at height can then be described by the following
probability distribution function:

( ) √

(2.1)

26/08/2020 ForsChem Research Reports Vol. 5, 2020-14 (4 / 22)


www.forschem.org
A Barometric Formula without the
Hydrostatic Pressure Assumption
Hugo Hernandez
ForsChem Research
hugo.hernandez@forschem.org

corresponding to the probability distribution of the absolute value of a normal random


variable, where is the standard deviation in vertical molecular speed at height .

The vertical dynamics of a single non-interacting molecule is presented in Figure 3, when solar
radiation absorption by the atmosphere is neglected. The molecular vertical speed at the
surface is , being the same in both upward and downward directions. As the molecule rises,
its speed decreases by the decelerating effect of gravity ( ), until the molecule completely
stops at a height . Then, the molecule falls increasing its speed as a result of gravity, until it
reaches the ground again. The same cycle is repeated at steady-state.

Figure 3. Vertical dynamic behavior of a single molecule. Molecular ground speed is . The
maximum height reached by the molecule is .

The vertical speed of the molecule ( ) can be described mathematically as a function of height
( ) using the following expression:

( ) √
(2.2)
where

(2.3)

3. Vertical Atmospheric Profiles

At steady-state, the number of molecules of a certain species crossing any horizontal plane at
altitude in the upward direction must be exactly the same as the number of molecules of
species crossing the plane in the downward direction. Mathematically, this can be expressed
as:

26/08/2020 ForsChem Research Reports Vol. 5, 2020-14 (5 / 22)


www.forschem.org
A Barometric Formula without the
Hydrostatic Pressure Assumption
Hugo Hernandez
ForsChem Research
hugo.hernandez@forschem.org

( ) ( ) ( )
(3.1)

where represents the molecular flux of species , and the superscripts and represent the
upward and downward direction, respectively.

On the other hand, a balance of momentum for each component in a horizontal atmospheric
section of volume at steady-state yields:

( ) ( ) ( ) ( ) ( ) ( )
( ) ( ) ̃( )
(3.2)
which can be transformed into :

( )( ( ) ( )) ( )( ( ) ( )) ̃( )
(3.3)

Assuming a symmetrical velocity distribution (particularly the normal distribution described in


Section 2) then:
( ) ( )
(3.4)
and considering positive the upward direction Eq. (3.3) becomes:

( ) ( ) ( ) ( ) ̃( )
(3.5)
where is the mean molecular speed in a single direction. Thus, for an infinitesimally small
height :

( ) ̃

(3.6)

Please notice that Eq. (3.6), obtained from the conservation of momentum, is conceptually the
correct expression to be used instead of the hydrostatic assumption (Eq. 1.1). Of course, the
equivalent pressure acting on the horizontal plane is related to , but it is not exactly
identical [8].

In addition, a molecular flux corresponds to the product between the molecular density ( ̃)and
the average molecular speed ( ) [8]:

( ) ̃( ) ( )
(3.7)

26/08/2020 ForsChem Research Reports Vol. 5, 2020-14 (6 / 22)


www.forschem.org
A Barometric Formula without the
Hydrostatic Pressure Assumption
Hugo Hernandez
ForsChem Research
hugo.hernandez@forschem.org

Therefore:
(̃ ) ̃ ̃
̃
(3.8)
or equivalently,
̃ ̃ ̃

(3.9)

The average molecular speed can be obtained from the vertical speed distribution (Eq. 2.1):

( ) ∫ ( ) ∫ √ √

(3.10)

However, due to spatial and temporal limitations, infinite molecular velocities are not possible.
Thus, we would expect that the mean molecular velocity should be lower than this ideal value.
Using a correction factor incorporating these limitations we obtain:

( ) √

(3.11)

Now, assuming that at steady-state the thermal energy of the molecules at a certain height is
homogeneously distributed in all directions, then:

( )

(3.12)
Therefore,
( )
( ) √

(3.13)
and

(3.14)
From which Eq. (3.9) becomes:

26/08/2020 ForsChem Research Reports Vol. 5, 2020-14 (7 / 22)


www.forschem.org
A Barometric Formula without the
Hydrostatic Pressure Assumption
Hugo Hernandez
ForsChem Research
hugo.hernandez@forschem.org

̃ ̃ ̃

(3.15)
Separating variables results in:
̃
( )
̃
(3.16)
which can be integrated into:

̃( )
( ) ∫ ∫ ∫ ( )
̃ ( ) ( )
(3.17)
where ̃ is the molecular density of species at a reference altitude (e.g. sea level).

Clearly, the molecular density profile can only be obtained if the atmospheric temperature
profile is known. Let us recall that the temperature profile is greatly influenced by solar
radiation absorption. However, if this effect is neglected (which is a common assumption for
tropospheric models), then temperature changes are only attributed to gravitational effects.

Considering the non-interacting model previously described, and assuming ideal gas molecules,
then the average kinetic energy of a molecule of species as a function of height will be:

〈 ( )〉
( )
(3.18)
where represents the degrees of freedom of motion of the molecular species .

Using Eq. (2.2), we obtain:

〈 〉
( ) ( ) ( )

(3.19)

Now, since the temperature of air at a certain height can be estimated as a function of the
average kinetic energy of the diatomic molecules, then:

( ) ( )
( ) ∑ ( ) ∑( ) ( )

(3.20)

where ( ) for , and ( ) is the molar fraction of species at height ,


given by:

26/08/2020 ForsChem Research Reports Vol. 5, 2020-14 (8 / 22)


www.forschem.org
A Barometric Formula without the
Hydrostatic Pressure Assumption
Hugo Hernandez
ForsChem Research
hugo.hernandez@forschem.org

̃( )
( )
̃( )
(3.21)
and

̃( ) ∑̃( )

(3.22)
where is the number of components in atmospheric air.

Therefore, the vertical temperature profile becomes:

( )
( ) ∑( )̃( )
̃( )
(3.23)
where the temperature contribution becomes zero when .

Assuming air as a pure system composed of molecules with an average molecular mass with
an average degrees of freedom of motion , then the temperature profile simply becomes:

( ) ( ) ( )

(3.24)
where is the constant lapse rate, estimated in this case as:

(3.25)
For this linear temperature profile,

(3.26)
And therefore Eq. (3.17) becomes:

̃( ) ( )
( ) ∫ ( )
̃ ( )
(3.27)
Integrating with respect to the altitude results in:

̃( ) ( )
( ) ( ) ( )
̃
(3.28)

26/08/2020 ForsChem Research Reports Vol. 5, 2020-14 (9 / 22)


www.forschem.org
A Barometric Formula without the
Hydrostatic Pressure Assumption
Hugo Hernandez
ForsChem Research
hugo.hernandez@forschem.org

And solving out for the molecular density we get:

( )
̃( ) ̃ ( )

(3.29)

Finally, the barometric formula of atmospheric air, under the steady-state assumption, can
simply be obtained from the ideal gas equation:

( ) ̃( ) ( )
(3.30)
where ( ) represents the equivalent pressure of the atmosphere at altitude .

Thus,
( )
( ) ( )

(3.31)

When a multicomponent model is considered, the molecular density of species is determined


by:
̃ ∫ ( )
̃( )
( )
(3.32)
where the temperature is given by Eq. (3.23).

These equations cannot be solved analytically, unless a certain assumption regarding the
temperature profile is done. For example, assuming a linear temperature profile:

( )
̃( ) ̃ ( )

(3.33)
where is the overall lapse rate approximately given by:

(3.34)

The previous equation is only an approximation because the atmospheric composition is


expected to change with altitude, and thus, the lapse rate is not truly constant.

26/08/2020 ForsChem Research Reports Vol. 5, 2020-14 (10 / 22)


www.forschem.org
A Barometric Formula without the
Hydrostatic Pressure Assumption
Hugo Hernandez
ForsChem Research
hugo.hernandez@forschem.org

The most general barometric expression considering multicomponent air and an arbitrary
temperature profile valid for an arbitrary altitude interval would then be:


( ) ∑ ( )

(3.35)

where the atmospheric conditions ( and ) at a certain altitude are known and used
as reference values.

In addition, the gravitational acceleration in this expression can be replaced by a suitable


altitude-dependent model.

4. Model Validation

Lente and Ősz [1] compared the different barometric formulas with the semi-empirical US
standard atmosphere (USSA) model [9]. In the USSA model, the temperature profiles are
represented by empirical piecewise models, whereas the pressure profile is obtained by
applying Pascal’s law in gases (Eq. 4 in [9]) using the empirical temperature, assuming ideal gas
behavior and standard sea level conditions ( )
[10].

The profiles described by the USSA model are compared in Figure 4 to the predictions of the
barometric formula derived in the previous Section (Eq. 3.31), using an estimated lapse rate
value (from Eq. 3.25) of , and the empirical lapse rate used by the USSA model
of . The standard deviation between the USSA model and the barometric formula
considering the theoretical estimation of the lapse rate and assuming was , with
an average relative absolute deviation of . The standard deviation between the USSA
model and the barometric formula using the empirical lapse rate and was and
the average relative absolute deviation of . Interestingly, when the theoretical lapse rate
is used with the present model the results are closer to the USSA model than when the
empirical lapse rate is considered.

The USSA model was used as a reference model. However, a model validation using
experimental data is also possible. For this purpose, a set of barometric pressure
measurements in the troposphere reported almost simultaneously from different locations
worldwide was collected from the following website: https://www.timeanddate.com/weather/.
The information was collected on June 9, 2020 between 18:00 and 19:00 GMT. The current
barometric pressures at 182 different locations with altitudes over 500 m above sea level are
presented in the Appendix 1. In addition, the standard sea level pressure and the barometric

26/08/2020 ForsChem Research Reports Vol. 5, 2020-14 (11 / 22)


www.forschem.org
A Barometric Formula without the
Hydrostatic Pressure Assumption
Hugo Hernandez
ForsChem Research
hugo.hernandez@forschem.org

pressure reported at Mt. Everest [11] were included in the dataset. The comparison of the
models with the collected set of experimental data is presented in Figure 5.

Figure 4. Barometric models comparison in the troposphere. Red line: US Standard Atmosphere
(USSA) model [9]. Blue line: New barometric model using Eq. (3.25) ( ) and
. Green line: New barometric model using the USSA empirical lapse rate ( )
and .

Figure 5. Barometric models validation using reported barometric data (gray circles) in the
troposphere. Red line: US Standard Atmosphere (USSA) model, . Blue line: New
barometric model using Eq. (3.25) ( ), . Green line: New
barometric model using the USSA empirical lapse rate ( ), .

We can observe that the experimental observations are somehow contained within the USSA
model ( ) and the completely theoretical barometric model without the

26/08/2020 ForsChem Research Reports Vol. 5, 2020-14 (12 / 22)


www.forschem.org
A Barometric Formula without the
Hydrostatic Pressure Assumption
Hugo Hernandez
ForsChem Research
hugo.hernandez@forschem.org

hydrostatic pressure assumption ( ). Considering the empirical lapse rate of


in the new model causes a deviation from the experimental observations particularly
at higher altitudes. This deviation does not penalize much the goodness of fit ( )
because only a few experimental observations are available at high altitudes.

Although the theoretical ideal gas model neglecting solar radiation does not accurately predict
the vertical pressure profile, it qualitatively explains why temperature decreases with altitude
in the troposphere. However, from a quantitative point of view, the effect of solar radiation is
of utmost importance. On one hand, the surface temperature increases, and energy can be
transferred as heat to the atmosphere; but on the other hand, the atmosphere itself also
absorbs solar radiation. The complexity of this situation increases by considering that solar
radiation intensity changes along the day, along the year, and is different at each location.
Furthermore, radiation absorption is greatly influenced by the local atmospheric composition.
However, as a result of the temporal and spatial averages in solar radiation and absorption, the
cooling effect caused by the gravitational force is partially compensated. That is why the
observed temperature change is smaller than the expected theoretical lapse rate when solar
radiation is neglected.

Figure 6. Vertical tropospheric temperature profiles. Blue diamonds: Experimental


observations. Red line: Theoretical ideal gas lapse rate without solar radiation (
). Green line: USSA model empirical lapse rate ( ). Blue line: Best
quadratic fit of experimental observations.

These differences might be illustrated by considering the vertical temperature profile models,
and comparing them with the experimental results. This comparison is presented in Figure 6.
Clearly, neither of the linear profiles correctly describes the experimental observations.
Particularly, a quadratic regression model provides a very good qualitative description of the

26/08/2020 ForsChem Research Reports Vol. 5, 2020-14 (13 / 22)


www.forschem.org
A Barometric Formula without the
Hydrostatic Pressure Assumption
Hugo Hernandez
ForsChem Research
hugo.hernandez@forschem.org

temperature profile. The empirical model obtained is the following (considering temperature in
and altitude in ):
( )
(4.1)
The empirical model presented in Eq. (4.1) can be approximated also as follows:

( )
(4.2)

Using this temperature model (Eq. 4.2), the vertical density and pressure profiles become (from
Eq. 3.32 and 3.35 assuming a single component):

̃ ( )
̃( )
(4.3)
( )
( )
(4.4)

Figure 7. Barometric models validation using reported barometric data (gray circles) in the
troposphere. Red line: US Standard Atmosphere (USSA) model, . Green line: New
barometric model using Eq. (4.4) ( ), . Green line: New barometric model
using Eq. (4.4) ( ), .

The vertical pressure profile given by Eq. (4.4) is comparatively shown in Figure 7. In addition,
the experimental data was used to fit the value using model (4.4). The result obtained

26/08/2020 ForsChem Research Reports Vol. 5, 2020-14 (14 / 22)


www.forschem.org
A Barometric Formula without the
Hydrostatic Pressure Assumption
Hugo Hernandez
ForsChem Research
hugo.hernandez@forschem.org

( ) represents approximately an equivalent maximum molecular velocity of


( )
√ .

The new barometric model presented in this work provides an improved description of the
barometric data ( ), while relying on valid first principles (mass,
momentum and energy balances) and not on the misconception of using Pascal’s law in gases.
The rate of change in pressure with respect to altitude can then be obtained for a linear
temperature profile from Eq. (3.31):

( )
̃ ̃
( )
(4.5)
and for Pascal’s law is:

̃
(4.6)

Clearly, the results obtained are very similar, but they are derived from different principles. This
explains why the hydrostatic pressure assumption, while being invalid for gases, can
satisfactorily describe the behavior of pressure in the atmosphere. Thus, it is not the purpose
to change the mathematical model but only to change its interpretation. Furthermore, the
larger rate of change in pressure compared to Pascal’s law is consistent with previous
theoretical developments [3,13] as well as experimental results [4].

5. Composition Profile

In addition to the pressure profile, the barometric model developed also allows estimating the
vertical composition profile of atmospheric air, as long as the temperature profile and the
ground-level composition are already known. For this purpose, the empirical temperature
profile considered by the USSA model [9] will be used. In addition, a representative
composition of dry unpolluted air at ground level is given in Table 1 [12].

Table 2 shows the composition of atmospheric air at different altitudes, as predicted by the
multicomponent model using the empirical lapse rate. Figure 8 illustrates the changes in air
composition with altitude, and Figure 9 shows the average molecular mass profile.

The multicomponent model predicts a composition decrease in all molecules heavier than the
average molecular mass (i.e. Oxygen, Argon, carbon dioxide, Krypton, nitrous oxide and
Xenon) and an increase in all molecules lighter than the average molecular mass of air (i.e.
Nitrogen, Neon, Helium, methane and Hydrogen). The average molecular mass of air is

26/08/2020 ForsChem Research Reports Vol. 5, 2020-14 (15 / 22)


www.forschem.org
A Barometric Formula without the
Hydrostatic Pressure Assumption
Hugo Hernandez
ForsChem Research
hugo.hernandez@forschem.org

observed in the troposphere to almost linearly decrease with altitude. The complete
atmospheric profiles obtained considering the empirical temperature profiles reported by the
USSA model are summarized in Appendix 2.

Table 1. Representative ground-level air composition [12] and additional air components
properties
Air component Molecular Degrees of %vol. / %mol
mass (g/mol) freedom
Nitrogen 28.01 5 78.084
Oxygen 32.00 5 20.946
Argon 39.95 3 0.934
Carbon Dioxide 44.01 9 0.036
Neon 20.20 3 1.82E-3%
Helium 4.00 3 5.24E-4%
Methane 16.04 15 1.60E-4%
Krypton 83.80 3 1.14E-4%
Hydrogen 2.02 5 5.00E-5%
Nitrous Oxide 44.01 9 3.00E-5%
Xenon 131.29 3 8.70E-6%
Average / Total 28.97 5 100

Table 2. Air composition at different altitudes obtained from the multicomponent model with
.
Altitude (km) 0 1 2 3 4 5 6 7 8 9 10 11
Composition
Nitrogen 78.08% 78.32% 78.56% 78.80% 79.05% 79.30% 79.55% 79.81% 80.07% 80.33% 80.60% 80.87%

Oxygen 20.95% 20.74% 20.54% 20.33% 20.11% 19.90% 19.67% 19.45% 19.22% 18.98% 18.75% 18.50%

Argon 9.34E-1% 9.02E-1% 8.70E-1% 8.38E-1% 8.07E-1% 7.76E-1% 7.46E-1% 7.16E-1% 6.86E-1% 6.57E-1% 6.28E-1% 6.00E-1%
Carbon
3.60E-2% 3.43E-2% 3.27E-2% 3.10E-2% 2.95E-2% 2.79E-2% 2.65E-2% 2.50E-2% 2.36E-2% 2.23E-2% 2.09E-2% 1.97E-2%
Dioxide
Neon 1.82E-3% 1.87E-3% 1.92E-3% 1.98E-3% 2.04E-3% 2.10E-3% 2.17E-3% 2.24E-3% 2.32E-3% 2.40E-3% 2.48E-3% 2.57E-3%

Helium 5.24E-4% 5.67E-4% 6.15E-4% 6.68E-4% 7.27E-4% 7.93E-4% 8.67E-4% 9.50E-4% 1.04E-3% 1.15E-3% 1.27E-3% 1.40E-3%

Methane 1.60E-4% 1.67E-4% 1.74E-4% 1.82E-4% 1.90E-4% 1.99E-4% 2.08E-4% 2.18E-4% 2.29E-4% 2.41E-4% 2.53E-4% 2.67E-4%

Krypton 1.14E-4% 9.56E-5% 7.98E-5% 6.64E-5% 5.49E-5% 4.52E-5% 3.71E-5% 3.02E-5% 2.45E-5% 1.97E-5% 1.58E-5% 1.26E-5%

Hydrogen 5.00E-5% 5.45E-5% 5.94E-5% 6.50E-5% 7.12E-5% 7.82E-5% 8.60E-5% 9.49E-5% 1.05E-4% 1.16E-4% 1.30E-4% 1.45E-4%

Nitrous Oxide 3.00E-5% 2.86E-5% 2.72E-5% 2.59E-5% 2.46E-5% 2.33E-5% 2.20E-5% 2.08E-5% 1.97E-5% 1.85E-5% 1.74E-5% 1.64E-5%

Xenon 8.70E-6% 6.26E-6% 4.47E-6% 3.16E-6% 2.22E-6% 1.55E-6% 1.07E-6% 7.28E-7% 4.92E-7% 3.29E-7% 2.17E-7% 1.42E-7%
Avg. Mol.
28.97 28.95 28.94 28.93 28.92 28.90 28.89 28.88 28.86 28.85 28.84 28.83
mass (g/mol)

26/08/2020 ForsChem Research Reports Vol. 5, 2020-14 (16 / 22)


www.forschem.org
A Barometric Formula without the
Hydrostatic Pressure Assumption
Hugo Hernandez
ForsChem Research
hugo.hernandez@forschem.org

Figure 8. Atmospheric air composition changes with altitude predicted by the multicomponent
model with . Left plot: Major air components. Right plot: Minor air components.

Figure 9. Vertical profile of the average molecular mass of atmospheric air, predicted by the
multicomponent model with .

6. Conclusion

In this report, a barometric formula was derived assuming a steady-state condition, and a
normal distribution of vertical molecular velocities at each altitude, by using only conservation
equations. In this derivation, Pascal’s law (Eq. 1.1), which is a valid law for dense fluids but
invalid for gases, was not used. Instead, Pascal’s law was substituted by an equation obtained
from the conservation of momentum considering a thin, horizontal section of atmospheric air
(Eq. 3.6). The solution of this equation requires knowledge of the vertical temperature profile,
and thus, the energy conservation equation needs to be solved first. While a simple
temperature model can be obtained by neglecting the effect of solar radiation absorption, such
model is far from reality. For this reason, and due to the complexity of the conservation of
energy considering solar radiation, an empirical vertical temperature profile is used instead.
Particularly, the empirical profiles reported in the US Standard Atmospheric model [9] are
used. A correction factor accounting for the limitations in molecular speeds was fitted using a

26/08/2020 ForsChem Research Reports Vol. 5, 2020-14 (17 / 22)


www.forschem.org
A Barometric Formula without the
Hydrostatic Pressure Assumption
Hugo Hernandez
ForsChem Research
hugo.hernandez@forschem.org

set of barometric pressure measurements reported almost simultaneously at different


locations worldwide. The correction factor thus obtained ( ) fitted very well
( ) the experimental data. Furthermore, using this numerical value for the
correction factor, the expression obtained closely resembles the formula obtained using
Pascal’s law, with an estimated pressure drop with altitude compared to the drop predicted by
Pascal’s law in the troposphere only . This may also explain why barometric models based
on the hydrostatic pressure assumption have been successfully used to describe the vertical
atmospheric pressure profile, despite the fact that Pascal’s law does not apply to gases.

Additional calculations using the same empirical temperature profile allow predicting the
changes in air composition with altitude. It is observed that the composition of light
components (below the molecular mass average) increase with increasing altitude, whereas
the composition of heavy components (above the molecular mass average) decrease with
altitude. For this reason, the average molecular mass of air decreases with altitude; particularly
in the troposphere, such decrease is almost linear.

By considering the particular temperature profile in any other atmospheric layer, Eq. (3.35) can
be used as the most general barometric formula to obtain the corresponding vertical pressure
profile.

Acknowledgments

The author gratefully acknowledges Prof. Jaime Aguirre (Universidad Nacional de Colombia)
for the several helpful discussions on the topic, and for reading the manuscript.

This research did not receive any specific grant from funding agencies in the public,
commercial, or not-for-profit sectors.

References

[1] Lente, G., & Ősz, K. (2020). Barometric formulas: various derivations and comparisons to
environmentally relevant observations. ChemTexts, 6, 1-14.

[2] Pascal, B. (1663). Traites de l'equilibre des liqueurs, et de la pesanteur de la masse de l'air.
Chez Guillaume Desprez, Paris.

[3] Hernandez, H. (2020). Pascal’s Law in Gases. ForsChem Research Reports, 5, 2020-09. doi:
10.13140/RG.2.2.36166.09285.

26/08/2020 ForsChem Research Reports Vol. 5, 2020-14 (18 / 22)


www.forschem.org
A Barometric Formula without the
Hydrostatic Pressure Assumption
Hugo Hernandez
ForsChem Research
hugo.hernandez@forschem.org

[4] Hernandez, H. (2020). Testing Pascal's Law in Gases using Free Fall Experiments. ForsChem
Research Reports, 5, 2020-12. doi: 10.13140/RG.2.2.35747.89120.

[5] Jordan, P. C. (1979). Chemical Kinetics and Transport. Plenum Press, New York.

[6] Maxwell, J. C. (1860). Illustrations of the dynamical theory of gases. - Part I. On the motions
and collisions of perfectly elastic spheres. The London, Edinburgh, and Dublin Philosophical
Magazine and Journal of Science, 19(124), 19-32.

[7] Boltzmann, L. (1872). Weitere Studien über das Wärmegleichgewicht unter Gasmolekülen.
K. Acad. Wiss.(Wein) Sitzb., II Abt, 66.

[8] Hernandez, H. (2019). Calculation of Molecular Fluxes and Equivalent Pressure in Ideal
Gases. ForsChem Research Reports, 4, 2019-03. doi: 10.13140/RG.2.2.35898.44483.

[9] NOAA, NASA & USAF (1976). US standard atmosphere, 1976. Washington.

[10] Corda, S. (2017). Introduction to aerospace engineering with a flight test perspective. John
Wiley & Sons. p. 186.

[11] West, J. B. (1999). Barometric pressures on Mt. Everest: new data and physiological
significance. Journal of Applied Physiology, 86(3), 1062-1066.

[12] Brimblecombe, P. (1996). Air Composition and Chemistry. 2nd Ed. Cambridge University
Press, New York. p. 2.

[13] Hernandez, H. (2020). Effect of External Forces on the Macroscopic Properties of Ideal
Gases. ForsChem Research Reports, 5, 2020-07. doi: 10.13140/RG.2.2.33193.21608.

Appendix 1. Worldwide Barometric Pressure Dataset

The data presented here corresponds to a set of barometric pressure measurements in the
lower layer of the atmosphere (troposphere), reported almost simultaneously from different
locations worldwide. The data was obtained from the following weather website:
https://www.timeanddate.com/weather/, the 9th of June 2020 between 18:00 and 19:00 GMT. A
total of 180 different locations with altitudes over 500 m above sea level were collected. In
addition, the standard sea level pressure and the barometric pressure reported at Mt. Everest
[11] were included as extreme values in the dataset.

26/08/2020 ForsChem Research Reports Vol. 5, 2020-14 (19 / 22)


www.forschem.org
A Barometric Formula without the
Hydrostatic Pressure Assumption
Hugo Hernandez
ForsChem Research
hugo.hernandez@forschem.org

Location Altitude (m) Pressure (Pa) Location Altitude (m) Pressure (Pa) Location Altitude (m) Pressure (Pa)
Abéché 540 95000 Edmonton 692 93400 Pedro Juan Caballero 662 94300
Abilene 524 95000 El Paso 1142 88900 Petrópolis 838 92200
Adama 1621 84700 Entebbe 1160 88900 Pirassununga 632 94800
Addis Ababa 2354 77800 Eskişehir 792 92300 Poprad 670 93600
Agadez 504 95000 Ferizaj 582 94300 Prince George 573 94500
Aguascalientes 1870 82700 Francistown 989 90900 Pristina 599 94100
Alajuela 955 91200 Freistadt 565 94600 Provo 1386 87000
Albuquerque 1510 85100 Fresnillo 2187 79700 Puebla 2155 79100
Ali Sabieh 713 92400 Ghardaïa 522 95700 Pune 567 94300
Alice Springs 581 95400 Gitarama 1872 82000 Puyo 950 92100
Almaty 790 92600 Granada 698 93800 Pyatigorsk 533 95200
Amarillo 1117 89000 Guadalajara 1543 84400 Querétaro 1817 83200
Amman 816 92300 Guarulhos 759 93400 Quito 2826 73800
Ankara 871 91600 Guatemala City 1533 86100 Rapid City 977 90400
Antananarivo 1287 87800 Guiyang 1067 89500 Regina 577 94700
Arequipa 2345 78300 Helena 1239 88100 Riyadh 590 94400
Arusha 1366 87000 Hesperia 971 91300 Rustenburg 1166 89200
Aurora 1647 83300 Hovd 1396 85600 Rwamagana 1543 85200
Bandung 709 93700 Hyderabad 507 95100 Sahuarita 824 92400
Bangalore 876 91600 Ibarra 2225 79400 Saint Thomas 760 93000
Baotou 1069 89000 Indore 554 93900 Salamanca 802 92700
Barquisimeto 569 95100 Innsbruck 570 94500 Salt Lake City 1312 87700
Béchar 779 92900 Islamabad 554 94400 Salta 1184 88300
Belo Horizonte 860 92500 Jaén 565 95300 San José 1161 89000
Bern 533 95100 Jerusalem 769 92500 San Luis Potosí 1876 81200
Bethlehem 753 92600 Kaduna 607 94600 San Marino 666 93400
Billings 946 91100 Kampala 1230 88200 San Salvador 667 94200
Bishkek 760 92900 Kathmandu 1305 87200 Santa Ana 650 94400
Bismarck 513 95200 Kempten 673 93400 Santa Fe 2132 79000
Bitola 615 93800 Kigali 1555 85100 Santiago 576 95000
Bloemfontein 1397 86500 Kiruna 532 95700 Sao Paulo 765 93300
Bogotá 2618 75700 Köniz 579 94600 Sarajevo 526 95000
Boise 822 92900 Kunming 1909 81600 Schwyz 515 95300
Brasilia 1091 90200 Lakewood 1683 83100 Sea Level [9] 0 101325
Bucaramanga 956 91100 Lanchow 1531 85000 Sion 514 95100
Butare 1750 83200 Las Vegas 629 95000 Sofia 551 94600
Calgary 1063 89500 León 1808 82000 Stavropol 580 94600
Cali 955 91200 Lhasa 3654 64300 Tabora 1207 88600
Campinas 686 94100 Lilongwe 1031 90500 Taiyuan 790 92400
Canberra 577 96000 Machu Picchu 2074 81100 Tegucigalpa 1054 90300
Caracas 875 91700 Madrid 658 94300 Tehran 1179 88900
Carson City 1425 86500 Malatya 960 90700 Tepic 936 90900
Cheyenne 1855 81200 Manizales 2139 79800 Texcoco 2250 77800
Chihuahua 1432 85500 Maun 943 91400 Thimphu 2307 75800
Chita 666 92800 Medellín 1479 86400 Tskhinvali 877 91700
Chur 592 94200 Medina 606 94500 Tucson 758 93100
Ciudad Juárez 1130 88700 Mendoza 767 92700 Tuxtla Gutierrez 536 95200
Cochabamba 2584 76200 Mexico City 2241 77900 Ulaanbaatar 1298 86100
Constantine 557 95300 Midland 847 92200 Ürümqi 832 91600
Cranbrook 933 91300 Molepolole 1145 89500 Valladolid 712 93700
Cuenca 2554 76200 Monterrey 545 95000 Valverde 571 95400
Cuernavaca 1522 85700 Mt. Everest [11] 8848 33600 Victorville 830 92800
Curitiba 915 91500 Munich 526 95200 Whistler 674 93800
Damascus 689 93400 Mwanza 1146 89000 Whitehorse 640 93400
Denver 1593 83800 Nairobi 1680 83700 Windhoek 1680 84000
Dire Dawa 1204 87800 Nekemte 2110 79700 Yaoundé 717 93500
Diyarbakir 671 93500 Nikšić 635 93600 Yeghegnadzor 1223 87400
Dodoma 1125 89400 Oaxaca 1566 84400 Yerevan 997 89800
Durango 1880 81500 Ogden 1310 87700 Zacatecas 2440 77400
Dushanbe 832 92100 Ouarzazate 1151 89200 Zaria 659 94100
Ebebiyín 564 94800 Oyem 664 93800
Ecatepec 2250 77800 Paradise 643 94800

26/08/2020 ForsChem Research Reports Vol. 5, 2020-14 (20 / 22)


www.forschem.org
A Barometric Formula without the
Hydrostatic Pressure Assumption
Hugo Hernandez
ForsChem Research
hugo.hernandez@forschem.org

Appendix 2. Complete Atmospheric Profiles

The following figures summarize the atmospheric vertical profiles (up to 86 km) obtained with
the barometric formula developed in this work. The starting point is the empirical temperature
profile presented in the 1976 U.S. Standard Atmosphere model [9].

Figure A1. Empirical vertical profile of the atmospheric air temperature, up to .

Figure A2. Vertical profile of the atmospheric air molecular density, up to , calculated
from Eq. (3.31). Left plot: Molecular density in original scale. Right plot: Molecular density in
logarithmic scale.

26/08/2020 ForsChem Research Reports Vol. 5, 2020-14 (21 / 22)


www.forschem.org
A Barometric Formula without the
Hydrostatic Pressure Assumption
Hugo Hernandez
ForsChem Research
hugo.hernandez@forschem.org

Figure A3. Vertical profile of the atmospheric air pressure, up to , calculated from Eq.
(3.34). Left plot: Pressure in original scale. Right plot: Pressure in logarithmic scale.

Figure A4. Vertical profile of the atmospheric air pressure, up to , calculated from the
multicomponent barometric model. Left plot: Pressure in original scale. Right plot: Pressure in
logarithmic scale.

Figure A5. Vertical profile of the atmospheric air composition, up to , calculated from the
multicomponent barometric model. Left plot: Major components (composition in % mole).
Right plot: Minor components (composition in logarithmic scale of molar fraction).

26/08/2020 ForsChem Research Reports Vol. 5, 2020-14 (22 / 22)


www.forschem.org

You might also like