ArticlePDF Available

Thermal Enhancement on Planetary Bodies and the Relevance of the Molar Mass Version of the Ideal Gas Law to the Null Hypothesis of Climate Change

Authors:

Abstract and Figures

Presented here is a simple and reliable method of accurately calculating the average near surface atmospheric temperature on all planetary bodies which possess a surface atmospheric pressure of over 0.69kPa, by the use of the molar mass version of the ideal gas law. This method requires a gas constant and the near-surface averages of only three gas parameters; the atmospheric pressure, the atmospheric density and the mean molar mass. The accuracy of this method proves that all information on the effective plus the residual near-surface atmospheric temperature on planetary bodies with thick atmospheres, is automatically 'baked-in' to the three mentioned gas parameters. It is also known that whenever an atmospheric pressure exceeds 10kPa, convection and other modes of energy transfer will totally dominate over radiative interactions in the transfer of energy, and that a rising thermal gradient always forms from that level. This rising thermal gradient continues down to the surface, and even below it if there is a depression or a mine-shaft present. This measured thermodynamic situation, coupled with other empirical science presented herein, mean that it is very likely that no one gas has an anomalous effect on atmospheric temperatures that is significantly more than any other gas. In short; there is unlikely to be any significant net warming from the greenhouse effect on any planetary body in the parts of atmospheres which are >10kPa. Instead, it is proposed that the residual temperature difference between the effective temperature and the measured near-surface temperature, is a thermal enhancement caused by gravitationally-induced adiabatic auto-compression, powered by convection. A new null hypothesis of global warming or climate change is therefore proposed and argued for; one which does not include any anomalous or net warming from greenhouse gases in the tropospheric atmospheres of any planetary body.
Content may be subject to copyright.
Earth Science
s
2018; 7(3): 107-123
http://www.sciencepublishinggroup.com/j/earth
doi: 10.11648/j.earth.20180703.13
ISSN: 2328-5974 (Print); ISSN: 2328-5982 (Online)
Thermal Enhancement on Planetary Bodies and the
Relevance of the Molar Mass Version of the Ideal Gas Law
to the Null Hypothesis of Climate Change
Robert Ian Holmes
Science & Engineering Faculty, Federation University, Ballarat, Australia
Email address:
To cite this article:
Robert Ian Holmes. Thermal Enhancement on Planetary Bodies and the Relevance of the Molar Mass Version of the Ideal Gas Law to the
Null Hypothesis of Climate Change. Earth Sciences. Vol. 7, No. 3, 2018, pp. 107-123. doi: 10.11648/j.earth.20180703.13
Received: February 25, 2018; Accepted: March 14, 2018; Published: April 13, 2018
Abstract:
Presented here is a simple and reliable method of accurately calculating the average near surface atmospheric
temperature on all planetary bodies which possess a surface atmospheric pressure of over 0.69kPa, by the use of the molar
mass version of the ideal gas law. This method requires a gas constant and the near-surface averages of only three gas
parameters; the atmospheric pressure, the atmospheric density and the mean molar mass. The accuracy of this method proves
that all information on the effective plus the residual near-surface atmospheric temperature on planetary bodies with thick
atmospheres, is automatically ‘baked-in’ to the three mentioned gas parameters. It is also known that whenever an atmospheric
pressure exceeds 10kPa, convection and other modes of energy transfer will totally dominate over radiative interactions in the
transfer of energy, and that a rising thermal gradient always forms from that level. This rising thermal gradient continues down
to the surface, and even below it if there is a depression or a mine-shaft present. This measured thermodynamic situation,
coupled with other empirical science presented herein, mean that it is very likely that no one gas has an anomalous effect on
atmospheric temperatures that is significantly more than any other gas. In short; there is unlikely to be any significant net
warming from the greenhouse effect on any planetary body in the parts of atmospheres which are >10kPa. Instead, it is
proposed that the residual temperature difference between the effective temperature and the measured near-surface
temperature, is a thermal enhancement caused by gravitationally-induced adiabatic auto-compression, powered by convection.
A new null hypothesis of global warming or climate change is therefore proposed and argued for; one which does not include
any anomalous or net warming from greenhouse gases in the tropospheric atmospheres of any planetary body.
Keywords:
Climate Sensitivity, Greenhouse Effect, Global Climate Change, Global Warming, Earth Temperature,
Venus Temperature, Auto-Compression, Atmospheric Thermal Gradient
1. Introduction
The basis of this work was first published in 2017 [23].
Here is presented a more comprehensive version, which
includes the following; a discussion on the causes of the late
20
th
century warming, newly published papers, a re-
assessment of the accuracy of the temperature of Mars, a
detailed discussion about Venus and an outline of the
problems with the currently accepted ‘null’ hypothesis of
climate. A new ‘null’ hypothesis of climate is advanced in
this work, which excludes any significant anomalous
warming effects arising from atmospheric greenhouse gases.
It will be shown that any anomalous warming effects of
greenhouse gases (GHG) such as CO
2
, are likely subjected to
a 100% rate of negative feedback in all troposphere’s, and
that this appears to be inherent to all planetary atmospheric
systems. The fine detail of these feedbacks will not be
outlined here, but the open nature of the atmosphere, coupled
with the following thermodynamic, scientific and other
arguments, indicate that this new null hypothesis of climate is
needed and fully makes sense. The present ‘null hypothesis’
of climate assumes - without empirical evidence - that there
is a tropospheric greenhouse effect (GHE); meaning an
anomalous net warming from greenhouse gases like CO
2
.
This effect supposedly causes significant net warming in the
troposphere [24] - even though this hypothetical warming has
Earth Sciences 2018; 7(3): 107-123 108
never actually been empirically measured, quantified and
then attributed to GHG in any published, peer-reviewed
scientific study to date.
1.1. Thermal Gradients Appear in All Atmospheres Above a
Pressure of 10kPa
It is known that planetary bodies which have thick
atmospheres, naturally set up a rising thermal gradient in that
part of the atmosphere which is higher than a pressure of
10kPa, until that bodies’ surface is reached [1] (Figure 1). Less
well known is that this rising temperature gradient continues
even below the surface [2] making it problematic to attribute
this thermal gradient to the GHE. In this denser part of the
atmosphere, the troposphere, convection and adiabatic auto-
compression effects rule over radiative or ‘greenhouse’ effects
[21] in energy transfers, in the determination of atmospheric
temperatures and in the formation of the thermal gradient.
However, higher up in the atmosphere, once the atmospheric
pressure drops below 10kPa then radiative effects dominate
energy transfers. This is because the atmosphere there is too
thin to initiate convection or any warming due to auto-
compression. Although the term ‘auto-compression’ may be
unfamiliar to some, this can be seen as simply an engineering
term for what meteorologists call the ‘lapse rate’ and
astronomers call the ‘Kelvin-Helmholtz’ contraction. Under
the latter, the contraction and compression of an inter-stellar
molecular gas cloud under gravity, achieves such enormously
high temperatures that nuclear fusion initiates, and a star is
born [3]. Support for the idea of a permanent thermal gradient
caused by the action of gravity on a thick atmosphere, in the
presence of the solar flux comes from recent work by Nikolov
& Zeller [4].
Figure 1. A thermal gradient appears in all planetary atmospheres >10kPa [1].
Using this knowledge, an exacting yet simple method is
introduced, which enables the average near-surface
atmospheric temperature of any planetary body with an
atmospheric pressure of over 10kPa, to become easily and
quickly calculated. A molar mass version of the ideal gas law
is utilised (formulas 5 and 6), which consists of one fixed gas
constant and three basic atmospheric gas parameters; the
average near-surface atmospheric pressure, the average near-
surface atmospheric density and the mean molar mass of the
near-surface atmosphere.
This formula proves itself here, to be not only more
accurate than any other method heretofore used but is far
simpler to calculate. It requires no input from parameters
previously thought to be essential for the calculation of
atmospheric temperatures, such as; solar insolation, albedo,
greenhouse gas content, ocean circulation and cloud cover
among many others. The reason these are not required, is
because they, (and all others) are already automatically
‘baked-in’ to the three gas parameters mentioned. Note that
although terms for insolation intensity and auto-compression
are not used in the formula, it is proposed that these two are
still what virtually determine an average near-surface
planetary atmospheric temperature.
1.2. Venus: The Planet Which Is Hard to Explain Using the
Greenhouse Effect of CO
2
There has always been difficulty in explaining, or in
formulating a simple method to satisfactorily explain or
calculate the very high surface atmospheric temperature of
the planet Venus using conventional mathematical means or
by employing the greenhouse gas hypothesis. Here, the molar
mass version of the ideal gas law will be used to simply and
accurately determine the surface temperature of this planet,
by the measurement of three gas parameters and the
knowledge of one fixed gas constant.
Whatever hypothesis is used to explain the Earth’s
temperature, it must also take into account the universality of
the physical laws of nature and of thermodynamics. For
instance, it must explain how a universal atmospheric
thermal gradient and enhancement [1] that is widely
attributed to the action of a wholly above-surface GHE, can
still continue on with its gradient unchanged, to below the
surface level as it does in a mine-shaft [2]. And how this
same gradient/enhancement appears in atmospheres with
109 Robert Ian Holmes: Thermal Enhancement on Planetary Bodies and the Relevance of the Molar Mass
Version of the Ideal Gas Law to the Null Hypothesis of Climate Change
virtually no greenhouse gases present. And must also explain
how the temperature in the Venusian atmosphere, at the same
pressure as the Earth’s surface, relates exactly to the Earth’s
average surface temperature once the different levels of solar
insolation [5] are taken into account - despite the large
differences in atmospheric greenhouse gas content. The
Venusian lapse rate, perhaps surprisingly, is very similar to
Earth’s at 7.7 K/km but extends much higher, to at least
50km [6]. A little below that height at 49km is where a
pressure of 1atm is to be found and is where a temperature of
~340K has been measured [6,78] to prevail.
Table 1. Earth’s average temperature at 1atm vs the Venusian temperature at the same pressure.
Planet Temperature 1atm Relative Solar Insolation Fourth Root Comparison Temperature
Earth 288 Kelvin 1.00 1.000 288 Kelvin
Venus 340 Kelvin 1.91 1.176 289 Kelvin
The temperature of a planetary body in space varies with
the fourth-root of the power incident upon it, meaning that
the temperature of Venus at 1atm (Tv) should be the fourth-
root of 1.91 times the temperature on Earth at 1atm (Te).
Venus receives 1.91 times the solar insolation of Earth [5]
(Table 1).
Tv = 1.91
x Te (1)
The temperature in the Venusian atmosphere from
Venera’s 8, 9, 10, 11 and 12 and from the Pioneer Sounder at
1atm, averages ~340K [58] [59]. This average temperature,
divided by the fourth-root of the insolation difference, results
in 289K - a value very close to Earth’s average surface
temperature at 1atm. Yet Venus has a 96.5% greenhouse gas
atmosphere, compared to Earth’s at just 2.5% [53]. It’s hard
to imagine atmospheres with such a differing greenhouse gas
content, yet there still remain very strong similarities in the
lapse rate, in the rate of the thermal gradient and as seen here,
in the relative insolation-adjusted temperatures at 1atm.
These measurements, relationships and the similarity of the
thermal gradients point strongly towards the existence of a
universal physical law which governs planetary atmospheric
temperatures - and one which does not take into account the
relative greenhouse gas contents; instead, this law clearly
operates as if GHG are not special.
It is now possible to solve for Venus at 1atm for density,
thus;
ρ = PM/RT (2)
ρ = 101.3 x 43.45 / 8.314 x 340 = 1.556kg/m³
The density at 1atm, assuming the atmosphere remains
well mixed at a height of 49km, calculates out at 1.556kg/m³.
Here the pressure has been chosen, and almost certainly the
molar mass remains the same as the surface, therefore this
has isolated any changes in the only free parameter
remaining; the density. The differences to Earth’s surface
parameters are now clear and are caused by the pressure of a
heavy dense atmosphere, offset by a density increase
mitigated by the higher insolation. The comparison result is a
50% rise in molar mass, which in isolation would translate
into a strong warming from 288K to 432K. But there is also a
27% rise in density over Earth at 1atm, which relates to
considerable offsetting cooling, resulting in the final
temperature of 340K. Revealed here is the advantage of
choosing a familiar pressure to work from rather than the
unfamiliar Venusian surface pressures. It can be seen that the
measured temperature difference from 288K to 340K is
directly related to the 50% higher molar mass of the
atmosphere combined with the 27% higher insolation-
moderated atmospheric density and very likely not to its
enhanced greenhouse properties.
Looking at the surface parameters on Venus can also be
instructive; again, there is the same 50% increase over Earth
in molar mass, which in isolation brings the initial base
warming to 432K. Then there remains the (very familiar on
Earth) battle between pressure and density, which finally
determines temperature. Here pressure clearly wins out with
a surface pressure that is 91x Earth’s, and density settles at
53x Earth’s. Do these numbers point to the end result of a
‘runaway greenhouse effect’, or to just what would be
expected from gas thermodynamics? Taking into account all
factors, the evidence suggests the latter. What very likely
determines these final numbers is the relationship between
the enormous mass of the Venusian atmosphere, the auto-
compression and the energy put into the upper atmosphere by
the Sun. There are many reasons to conclude that there is no
net warming from the Venusian CO
2
.
Of note is that very little or no direct solar insolation
reaches the Venusian surface [58, 59], certainly no more than
10% of that which reaches the Earth’s surface. In addition,
the lowest several kilometres of the Venus atmosphere are
not a gas, but a super-critical fluid. The critical pressure of
CO
2
is 7,380 kPa and the critical temperature is +30°C, so
the conditions on the near-surface of Venus dictate that the
entire atmospheric surface layer, to a depth of approximately
~4km must be a super-critical fluid. Five problems can be
readily identified with regard to the possibility that the GHE
of CO
2
is the cause of Venus’s high surface temperatures, as
is currently claimed by NASA, the IPCC and most
mainstream climate scientists;
1) The first question that might be asked is; can a highly
compressed and super-heated super-critical fluid that is
more like an ocean than a gas, still possess the
greenhouse properties of an ordinary atmospheric gas?
This seems to be highly unlikely. However, it is true that
fermions, (of which CO
2
is made) when highly
compressed, increase the width of their
absorption/emission bands, (because the Pauli Exclusion
Principle [72] prevents fermions from being in the same
state and in the same place.) Whether this factor has
Earth Sciences 2018; 7(3): 107-123 110
affected the surface super-critical fluid sufficiently to
create a gas-like GHE is unknown at present.
2) A second problem with regard to the GHE claim for
Venus is that the atmosphere is very thick and is
optically opaque more like a thick soup than
transparent like the Earth’s atmosphere is.
Measurements from the surface of Venus show that
<20W/m² of direct short-wave solar insolation [58, 77]
actually makes it to the surface of Venus to warm the
surface for the up-welling infra-red radiation to be
available to be captured by any possible atmospheric
GHE. In fact, direct solar insolation can be neglected
below a height of 60km, as virtually all direct solar
radiation below that level is ‘scattered’ by the thick
atmosphere. The flux of this scattered solar insolation
was measured on the surface by six separate landers and
appears to be very low [58] averaging <<10% of the
2,644 W/m² TOA insolation flux. In contrast, Earth
receives much more at 12% of its TOA insolation
directly onto the surface (161 W/m² of 1,366 W/m²)
[74] and much more if scattered, and atmospheric and
back-radiation are counted.
3) Third, Venus has a very slow rotation period, which
makes the Venusian ‘night’ ~58 Earth-days long [75].
During this long night, measurements have been taken of
the atmospheric and the surface temperatures, and they
remain basically the same all through the long night just
as they are during the long 58-day Venusian ‘day’. The
surface cools only very slightly from ~737K to ~732K
during this very long night. A question might reasonably
be asked here; “How can the GHE of CO
2
be responsible
for all this surface heat, by trapping upwelling longwave
radiation, emitted from absorbed direct solar insolation
and hence keeping the surface hot with re-emitted
downwelling radiation, when little or no direct Sun
arrives to the surface during the ‘day’ and when no Sun
at all arrives during the 58-day long ‘night’?”
4) Fourth, the very high albedo reduces Venus’s access to
solar insolation. Even though Venus’s TOA insolation is
~2x Earth’s, the reflectivity of Venus is so high at 75%
that this more than cancels out the higher TOA
insolation. This means that although it is closer to the
Sun, the Venusian atmosphere as a whole actually
absorbs much less Solar warmth than Earth does;
(2,644/4) x (1-0.75) = 165 W/m² vs (1,366/4) x (1-0.29)
= 242 W/m² for Earth. If Venus receives even less net
solar radiation than the Earth does, how can it maintain
a very much higher temperature profile in its
atmosphere because of this radiation?
5) Fifth, although as might be expected because of its high
density, the Venusian atmosphere moves only slowly at
the surface (<10km/hr), it rotates very rapidly at 70km
in height, the cloud tops level, circling the planet every
4 days at speeds of up to 100m/s (360km/hr) [76]. Why
does the Venusian atmosphere rotate westwards at sixty
times [73] the rotation speed of the planet, and what is
the mechanism driving and maintaining it? Given that
the atmosphere is open to space and can expand and
contract, and is in constant motion like this, how is the
GHE of CO
2
affected? Could it be subjected to
sufficient negative feedbacks to eliminate any net
warming from it altogether?
The Venus atmosphere is so hot that it radiates at the rate
of 15,000W/m² down to the surface, [79] even though less
than 20W/m² of direct solar insolation actually reaches the
surface. A conventional ‘GHE’ of the type described by the
IPCC is not possible with these numbers. If it is not the GHE,
then where does Venus get the vast amount of energy from to
keep such a heavy, thick atmosphere in motion and so very
hot? The answer proposed here is the same as for Earth; auto-
compression, adiabatic convection and the conversion of
higher-level atmospheric potential energy to lower-level
kinetic energy.
1.3. Molar Mass Version of Ideal Gas Law Accurately
Calculates Planetary Surface Temperatures
A version of the ideal gas law may be used to more
accurately determine surface temperatures of planets with
thick atmospheres than the S-B black body law, [7] if a
density term is added; and if kg/m³ is used for density instead
of gms/m³, the volume term V can be dropped. This formula
then may be known as the molar mass version of the ideal
gas law (Formulas 5 or 6).
The ideal gas law is; PV = nRT (3)
Convert to molar mass; PV = m/M.RT
Convert to density; PM/RT = m/V = ρ
Drop the volume term; ρ = P/(R.T/M) (4)
Find for temperature; T =
(
)
(5)
V = volume
m = mass
n = number of moles
T = near-surface atmospheric temperature in Kelvin
P = near-surface atmospheric pressure in kPa
R = gas constant (m³, kPa, kelvin¹, mol¹) = 8.314
ρ = near-surface atmospheric density in kg/m³
M = near-surface atmospheric mean molar mass (gm/mol
-1
)
Alternatively, the molar mass version of the ideal gas law
can be written thus;
T = PM/Rρ (6)
2. Methodology Involves Calculating the
Average Near-Surface Temperature of
Planets
Formula 5 is here used throughout:
Using the properties of Venus, [8]
111 Robert Ian Holmes: Thermal Enhancement on Planetary Bodies and the Relevance of the Molar Mass
Version of the Ideal Gas Law to the Null Hypothesis of Climate Change
 = 9200
(8.314x 65
43.45)
Venus calculated surface temperature = 739.7K
Using the properties of Earth from Wiki, [9]
 = 101.3
(8.314x 1.225
28.97)
Earth calculated surface temperature = 288.14K
Venus is calculated at 739.7K, which is given by NASA as
~740K. Earth is calculated at 288K, currently its quoted by
NASA [5] at 288K. It will be noted that the average
temperature of the surface of Titan was measured by the
Voyager 1, and by the Huygens lander [10] and was probably
used as an input to find the surface density; (the
independently-measured surface density on Titan could not
be found in the literature). The 94K will therefore come out
of the below formula, since it is a rearrangement of formula
3. This could be seen as a circular argument. However, it is
unlikely that if and when the density of Titan is directly
measured, for instance by the use of a dasymeter or similar, it
will be significantly different from the 5.25kg/m³ stated here.
Calculate for Titan, data [11];
 = 146.7
(8.314x 5.25
28.0)
Titan calculated surface temperature = 93.6K
Titan and Saturn share the same solar insolation, yet Titan is
much colder (39K colder) than Saturn, despite the moon having
8,000 times the concentration of the strong greenhouse gas
methane in its atmosphere than Earth does, or more relevantly, 3
times the methane concentration than Saturn does. Saturn has no
other significant GHG in its atmosphere. Titan is even colder at
1atm - which is the level the temperature is being measured on
Saturn. Why is Titan so cold, despite all the greenhouse gas it
possesses, and in spite of receiving just the same solar insolation
that Saturn does? Its density; at surface, the density is 27 times
that at 1atm on Saturn. Questions remain about where and how
Titan got such a dense and thick atmosphere, - and how the
small moon retains it. Also of interest is that as well as the 2.7%
Methane content, Titan has a 97% Nitrogen atmosphere, a gas
which has been claimed to take on some of the properties of a
greenhouse gas at the temperatures prevailing on Titan [80].
And like Venus, Titan’s atmosphere appears to be a ‘super-
rotator’ [81], meaning that the atmosphere rotates much faster
than the planetary surface does. Could this rapid motion be the
result of a negative feedback effect related to the almost 100%
of ‘greenhouse gases’ which comprise the atmospheres of both
planetary bodies?
Calculate for Earth’s South Pole, data [12];
 = 68.13
(8.314x 1.06
28.97)
Earth’s South Pole average calculated temperature = 224K
(-49°C)
It’s clear from these figures at the South Pole, that the low
temperature comes mainly from the low pressure.
Calculate for Mars [13];
 = 0.69
(8.314" 0.02
43.34)
 = 0.9
(8.314" 0.02
43.34)
Mars calculated surface temperature = 180K to 234K
The average temperature on Mars is reported as either
210K [13] or 191K [4]. As suspected from other work [1] [4]
this method of temperature calculation is tricky for Mars, due
to the very low and highly variable atmospheric pressure.
Pressures were measured [60] at the Viking 1 landing site and
varied between 0.69kPa and 0.9kPa, according to the season.
It is noted that it is only in atmospheres with a pressure of
over 10kPa that strong convection and a troposphere/tropopause
is formed, with its associated thermal gradient. Nevertheless, the
formula still provides a useful range of surface temperature by
the use of these lower pressures, which in fact extend evenly
across the measured actual. For Mars, the mid-point between the
summer and the winter pressures is used, which results in a
calculated temperature of 207K, compared to the measured
210K.
The gas giants will now be assessed; note that these
planets do not have a defined surface like the terrestrial
planets have, so here they are given a ‘surface’ by using the
Earth’s surface pressure of 101.3kPa (1atm) as a level to use
for this calculation.
Calculate for Jupiter [5];
 = 101.3
(8.314x 0.16
2.2 )
Jupiter calculated temperature at 1atm of pressure = 167K
Calculate for Saturn [5];
 = 101.3
(8.314x 0.19
2.07)
Saturn calculated temperature at 1atm of pressure =
132.8K
Calculate for Uranus [5];
 = 101.3
(8.314x 0.420
2.64 )
Uranus calculated temperature at 1atm of pressure = 76.6K
Calculate for Neptune [5];
 = 101.3
(8.314x 0.450
2.53 )
 = 101.3
(8.314x 0.450
2.69 )
For Neptune, NASA supplies two values for mean molar
mass; 2.53 and 2.69, this necessitated two separate
calculations to give a high and a low of calculated
Earth Sciences 2018; 7(3): 107-123 112
temperatures, as was done for Mars. Neptune’s calculated
temperature at 1atm gives a range of 68.5K to 72.8K. The
temperature on Neptune at 1atm of pressure is measured at
72K; this lies quite handily between the two calculated
temperatures. The calculated and actual average ‘surface’
temperatures of the eight planetary bodies are compared in
Table 2, along with the errors.
Table 2. Comparison of calculated and actual average ‘surface’ temperatures.
Planetary body Calculated temperature Kelvin Actual temperature Kelvin Error
Venus 739.7 740 0.04%
Earth 288.14 288 0.00%
South Pole of Earth 224 224.5 0.20%
Mars (low pressure) 180 to 234 averages 207 210 1.40%
Jupiter 167 165 1.20%
Saturn 132.8 134 0.89%
Titan 93.6 94 0.42%
Uranus 76.6 76 0.79%
Neptune 68.5 to 72.8 averages 70.7 72 1.90%
2.1. Explanation of the New ‘Null’ Hypothesis of Climate
Change Proposed Herein
The existing null hypothesis of climate change simply
assumes that there exists a 33°C ‘residual’ warming effect
[66], which in turn is assumed to be 100% produced by GHG
in the lower troposphere [24]. More assumptions are that
once CO
2
is emitted by humans to the atmosphere, it remains
there for ‘hundreds of years’ [24]; another assumption is that
the ice core record for CO
2
is correct (and not the plant
stomata record) and therefore that the CO
2
concentration in
1750 (the so-called pre-industrial level) was 280ppmv, and a
final assumption is that all of the increase from this assumed
280ppm to the measured present level is anthropogenic. It is
proposed here that most or all of these assumptions are
incorrect. The residence time for CO
2
is first shown to be
incorrect; it is in fact just 4yr [65,69,70,71].
The new ‘null’ hypothesis of climate change being put
forward here, is that in the case of Earth, solar insolation
provides the ‘first’ ~255*Kelvin – in accordance with the
black body law [14]; this being the ‘effective’ or the ‘base’
level. And a gravitationally induced thermal gradient caused
by auto-compression provides the ‘other ~33*Kelvin,
termed the ‘residual’, to arrive at the known and measured
average global temperature of 288Kelvin. The ‘residual’ is
not hypothesised to be provided by anomalous warming from
greenhouse gases, because if it was, it would not make sense
that the Venusian temperature at 1atm correlates exactly with
Earth’s temperature at the same pressure, when insolation
differences are allowed for.
And the consistent thermal gradients seen across
atmospheres and across all greenhouse levels above 10kPa
would not make sense, and the evidence presented in support
of the new null hypothesis as laid out in detail in section 3
would be violated. The result of the ‘thought experiment’
conducted in section 2.7-9 is also consistent with all these
findings. If the new null hypothesis of natural climate change
is to be violated, substantial and convincing empirical
scientific evidence would need to be brought to bear; it
would need to be in excess of that which has been presented
in the literature to date.
* These figures are disputed by recent work [4] however
this still would not change the conclusions here.
2.2. What Temperature Is and a Discussion About Maxwell
and Loschmidt’S Ideas
Temperature in a gas is a measure of the average kinetic
energy of the particles in the gas. When atmospheric gas
pressure exceeds 10kPa, a temperature gradient is set up from
that pressure level, [1] down to a planetary surface. This
thermal gradient constitutes a thermal enhancement and is
known and measured to continue even below the surface, if
there is for example, a mine shaft. It is hypothesised here,
that the cause of this thermal gradient is gravity-induced
auto-compression, and that along with insolation is an
essential part of the null hypothesis of climate. In general
terms, the surface temperature sets up convective overturning
of the troposphere, which is adiabatic through much of the
convection cycle [2], and this combines with gravitationally
induced atmospheric auto-compression to create the observed
tropospheric thermal enhancement and its associated
temperature gradient.
The origins of this thermal effect on gases go back to
James Maxwell, who, in his 1872 book ‘Theory of Heat’ [15]
demonstrated that the formation of the thermal gradient from
the tropopause downwards is assisted by convection and
more particularly, the increasing atmospheric pressure, which
itself is a result of a combination of the Earth’s gravitational
field and the atmospheric density.
“In the convective equilibrium of temperature, the
absolute temperature is proportional to the pressure..” James
Maxwell [15].
The idea of a thermal gradient naturally forming in any
column of gas in a gravitational field was first proposed in
the 1860’s by Loschmidt [16]. At the time, Maxwell thought
that this idea violated the second law of thermodynamics, yet
as has been shown here, derivations of Maxwell’s own ideal
gas law is an excellent predictor of temperatures whenever
an atmosphere is thick enough to be compressed in a
gravitational field.
The controversy between Loschmidt on one side, with
Maxwell and Boltzmann on the other, raged for some time
and was finally experimentally tested in 2007, with the
113 Robert Ian Holmes: Thermal Enhancement on Planetary Bodies and the Relevance of the Molar Mass
Version of the Ideal Gas Law to the Null Hypothesis of Climate Change
results published by Graeff [17] [56]. Graeff’s experiments
concluded that a gravitationally-induced temperature gradient
does spontaneously develop in sealed columns of both air
and water the bottom of the column being warmer than the
top. The theoretical amounts of warming according to Graeff
should be 0.07K/m and 0.04K/m respectively. Graeff’s
experimental apparatus reported 0.07K/m and 0.05K/m so
basically confirming Loschmidt’s predictions. The thermal
gradient appeared, despite the reverse gradient being
prevalent in the immediate environment of the experiment.
Loschmidt originally said that the second law of
thermodynamics needed to be re-stated to include the effects
of gravitational fields on fluids.
More recent work by Levy [57] has thrown some doubt on
Graeff’s work that a static column of a Maxwellian gas (such
as air) can spontaneously create a stable thermal gradient.
However, Levy’s work does show that a thermal gradient is
always created when a convective current, driven be a heat
source arises in a Maxwellian gas that is immersed in a
gravitational field; which is the exact case in all of the
planetary atmospheres examined here. Also revealed in the
work, is the theoretical basis for another observation - that
the rate of the thermal gradient is dependent on the average
molar mass of the atmosphere; with a higher molar mass
resulting in a stronger gradient. A gravity-induced thermal
gradient due to adiabatic compression is also found in the
deep ocean, commencing at 5,000m [57].
2.3. Auto-Compression Is Well Known and Used Daily in
Mining
Auto-compression is well known in underground mining
and is used by ventilation engineers to calculate how hot the
mine air will get, so that they know how much cooling air to
provide at each level. The effect of auto-compression can be
calculated by the following relationship;
Pe = Ps exp(gH/RT) (7)
Where;
Pe = absolute pressure at end of column (kPa)
Ps = absolute pressure at start of column (kPa)
g = acceleration due to gravity (m/s²)
H = vertical depth (m)
R = Standard Temperature (Kelvin)
T = Final Temperature (Kelvin)
As can be clearly seen, this effect primarily relies on
pressure and gravity, which will be different for each
planetary body.
2.4. Mechanism Is Adiabatic
Note that we are examining a largely adiabatic process
during convection. When a gas parcel expands adiabatically,
as it does when rising in a gravitational field, it does positive
work and the kinetic energy drops and so the temperature
drops. However, when a gas parcel is compressed, as it is
when it descends adiabatically in a gravitational field, then it
does negative work, and its kinetic energy rises and so its
temperature goes up. Why does the kinetic energy of the gas
rise when descending? It’s because some of its potential
energy is converted to enthalpy, so producing an increase in
pressure, specific internal energy and hence, temperature in
accordance with the following equation;
H = PV + U (8)
Where;
H = enthalpy (J/kg)
P = pressure (Pa)
V = specific volume (m³)
U = specific internal energy (kinetic energy)
2.5. Discussion on Maxwell vs Arrhenius and the
‘Greenhouse Effect’
Work in this area of gas physics was detailed in the 19
th
century. However, there is a strong difference between the
work and the views of the researchers Maxwell and
Arrhenius. Maxwell’s work [15] shows that temperatures in
the lower troposphere of Earth are primarily determined by
convection and the atmospheric mass/pressure/gravity
relationship. Arrhenius’s later work [18] completely ignored
this and determined that temperatures in the lower
troposphere of Earth are caused by the radiative effects of
greenhouse gases. There have been papers critical of
Arrhenius’s radiative effects ideas since 1909 [19]. Which
idea is correct is critical to the present, since if Arrhenius is
correct, then there should be some concern about CO
2
emissions, if the climate sensitivity is high enough. But if
Loschmidt’s version of Maxwell’s work is correct, then
doubling CO
2
will have no measurable effect on tropospheric
atmospheric temperatures, and the climate sensitivity will be
too low to be measurable.
What do atmospheric measurements actually show?
Measurements [20] of the effects of more CO
2
in the atmosphere
appear to strongly support Maxwell’s ideas. At pressures above
10kPa, “the extra CO
2
merely replaces water vapour” and little
difference is seen in temperatures – but at pressures below
10kPa more CO
2
is measured to cause strong cooling. One of
the main problems with the Arrhenius view, is that radiative
transfers are emphasised, and convection is virtually ignored as
a mode of heat transfer. Yet later work shows that not more than
11% of heat transfer in the troposphere is actually carried by
radiation [21]. Whether a small change in this already small
percentage can cause significant net warming in an open
atmosphere is highly debatable. A recent paper has supported
the Arrhenius view somewhat by quantifying a small forcing
due to increased atmospheric CO
2
[22], however, there has been
no confirmation of this in a follow-up paper. But there still
remains a lack of any paper in the literature, which quantifies
any warming that has been attributed to increasing atmospheric
CO
2
concentrations.
2.6. The Accuracy, Implications and Limitations of
Formulas 5 and 6
It is apparent that this simple formula calculates the near
Earth Sciences 2018; 7(3): 107-123 114
‘surface’ temperatures of many planetary bodies in our Solar
System very accurately (Figure 2). Specifically, all of those
which have atmospheres thick enough to form a troposphere
(i.e. possessing an atmospheric pressure of >10kPa). These
are; Venus, Earth, Jupiter, Saturn, Titan, Uranus and Neptune.
All calculated temperatures are within 1.2% of the NASA
reported ‘surface’ temperature (Mars’s temperature can also
be predicted but is <<10kPa, which is too low for convection
to occur). This accuracy is achieved without using the S-B
black body law, or the need to include terms for such
parameters as TSI levels, albedo, clouds, GHE or, for that
matter, adiabatic auto-compression.
All that is required to be able to accurately calculate the
average near-surface atmospheric temperature, is the relevant
gas constant and the knowledge of the three variable gas
parameters.
Figure 2. Actual temperature vs calculated temperature of 8 planetary bodies and the South Pole.
To be crystal clear about the limitations of the molar mass
version of the ideal gas law; the nature of the formula means
that it cannot in general be used to determine the cause of an
atmospheric warming or cooling event. However, by
isolating and examining changes to specific gas parameters,
it may be possible to determine what is not causing a specific
warming or cooling event - if the effect of the cause must
result in a large anomalous change in a specific gas
parameter or set of gas parameters.
2.7. A Thought Experiment Involving Two Planets
To more easily conceptualise and determine the source of
the thermal gradients and the surface temperature
enhancement which are known to exist on all planetary
bodies with thick atmospheres [1], the use of a thought
experiment is proposed.
Consider; two very Earth-like rocky planets with Earth-
like atmospheres orbiting at the same distance (1
astronomical unit) from the Sun. We provide one with an
atmosphere identical in every way to the present Earth’s,
containing 0.04% CO
2
by volume; let this planet be E1. Now
the other planet E2, is going to be identical in every way to
E1 except for a slight difference in the composition of the
atmosphere. E2 will be identical to E1’s atmosphere - but
with one important difference; it will contain twice the
concentration of CO
2
at 0.08% by volume.
Clearly the existing greenhouse gas hypothesis, and the
existing null hypothesis predicts that E2 should have a
significantly higher (~3K) surface temperature than E1,
because of its extra 0.04% of the greenhouse gas CO
2
[24].
This IPCC reports’ view is of a climate sensitivity at this
level, this is also backed by the ‘97% consensus’ [67].
Conversely, the new null hypothesis as presented here,
predicts that both planets will have virtually identical
temperatures. The dilemma is; how to determine which null
hypothesis is correct? This puzzle appears to be solvable in
the following manner.
How could a simple formula such as formula 5, which
contains no reference to the percentage of GHG in an
atmosphere, accurately predict the temperature of a planet
with a very specific percentage of GHG, such as planet E2?
Perhaps it would be informative to be aware of the wide
variation in the atmospheres of other planetary bodies - some
with up to 96% GHG in their atmospheres (Venus, Titan) -
and some others with virtually none (Jupiter, Saturn). A
simple formula with no reference to GHG in it would not be
expected to accurately predict the average atmospheric
temperature of eight such widely differing planetary
atmospheres, by the measurement of just three common
planetary gas parameters.
And yet it does (Table 2, Figure 2).
The only way that is possible, if the greenhouse gas
hypothesis is correct and these gases are special and cause
strong warming, is that changes in the greenhouse gases’
percentage in an atmosphere must alter the pressure and/or
density anomalously; - in such a way as to make formulae 5 fit.
2.8. The Search for an Anomalous Change in the Gas
Parameters*
If these two planets have almost the same
115 Robert Ian Holmes: Thermal Enhancement on Planetary Bodies and the Relevance of the Molar Mass
Version of the Ideal Gas Law to the Null Hypothesis of Climate Change
pressure/density/molar mass, and hence virtually the same
temperature, then the extra 0.04% of CO
2
must have had
almost no anomalous warming effect.
Figure 3. The present Earth (E1) is compared to a possible future Earth with a doubled CO
2
(E2).
*Postulate; the molar mass version of the ideal gas law is
correct.
1) For planet E2 to become 3°C warmer than E1, as is
claimed by the IPCC, the extra CO
2
must change one or
more of the three gas parameters very significantly and
anomalously - i.e. there has to be a large and an
anomalous effect on one or more of these three gas
parameters.
2) If the presence of the extra CO
2
does not change one or
more of the three gas parameters very significantly and
hence anomalously, then the greenhouse gas hypothesis
and the current null hypothesis must be incorrect and
are invalidated.
3) This can be regarded as a test of the here-presented
hypothesis that the present null hypothesis of climate
change is incorrect and needs to be changed to one in
which there is no net anomalous warming from GHG
such as CO
2
.
2.9. Assessing Whether an Anomalous Warming from a
Doubling of CO
2
Exists on Planet E2
Scenario 1: When the reality with regards to the
equivalence of the temperatures of Earth and Venus at the
same pressure is taken into account (see section 1.2). A
reasonable expectation would be that a 0.04% increase in
atmospheric CO
2
, which is a relatively heavy gas, could be
expected to result in the following approximate atmospheric
changes in the three gas parameters;
Pressure: an increase of 0.04%
Density: an increase of 0.05%
Molar Mass: an increase of 0.05%
Calculate using formula 5 a doubling of CO
2
from the
current level of 0.04%;
 = 101.34
8.314x 1.2256
28.984
Calculated temperature after doubling of CO
2
to 0.08%
288.25K
‘Reasonable Expectation’ equilibrium climate sensitivity
to CO
2
≈ 288.25 - 288.14 ≈ +0.11K
Under this circumstance, the climate sensitivity would in
fact be extremely small and difficult to estimate exactly, but
would be of the order +0.11°C. That is, twenty-seven times
smaller than the stated ‘likely’ climate sensitivity of 3°C
cited as the ‘median’ in the IPCC’s reports [24]. This
reasonable expectation of climate sensitivity to CO
2
of
+0.11°C is so low that it would be impossible to detect or
measure in the real atmosphere, even before any allowance is
made for the consumption of atmospheric O. But that small
number would likely be a maximum change, because if fossil
fuels are burned to create the emitted CO
2
then atmospheric
O will also be consumed, reducing that gas in the
atmosphere – and offsetting much of any net temperature
change that is generated by the extra CO
2
, reducing the
climate sensitivity further.
Scenario 2: The alternative scenario is that CO
2
possesses
the same net atmospheric warming properties ascribed to it
by the IPCC and most mainstream climate scientists. In this
case, E2 would have to become ~3°C warmer than E1, at
equilibrium [24]. What would need to change in the
parameters from scenario 1 - and by how much? And are
these changes reasonable / possible? First, there could not be
a change in the average molar mass, since the molar mass of
all the atmospheric constituents are known fully and are
known to always be well-mixed. This must remain at 28.984.
Therefore, by elimination, the entire 3°C of warming has to
come from an anomalous and significant change in either the
pressure, or the density, or both.
It is known from the formula, that when isolated, only a
decrease in density can cause a warming; an increase will cause
cooling. Similarly, when isolated, only an increase in pressure
causes warming; a decrease causes cooling. And so here is a
situation where a heavy gas is being added to the atmosphere;
thus one logically expected outcome may be a higher
atmospheric pressure, and so a higher temperature - and yet a
Earth Sciences 2018; 7(3): 107-123 116
higher pressure and a heavy gas would also surely indicate a
greater density, - and so a counter-balancing lowering of
temperature! This is a conundrum. How can this be resolved?
Perhaps the pressure and density should be taken separately, to
assist in clarifying the extent of the GHG anomaly problem.
If the 3°C of warming is not thought to be even partly a
result of density decreasing (which seems logical when a
heavy gas is being introduced) then what increase in pressure
would be required to explain all of the temperature change, if
density were to be held static? This scenario is shown in E2,
Figure 3.
Pressure: an increase of 1.00% due to greenhouse warming
Density: no anomalous change assumed
Molar Mass: no anomalous change possible
Calculate for a doubling of CO
2
from the current level of
0.04% (by volume);
 = 102.35
(8.314x 1.2256
28.984)
Calculated temperature after doubling of CO
2
to 0.08%
291.14K
To reach the required temperature here, the pressure would
have to increase anomalously by 1.00%. How likely is it that
the action of 0.04% more CO
2
increases the pressure at
equilibrium by twenty-five times through the GHE, over
what would be expected from its physical presence alone?
And this has to be seen as a minimum, because if the
pressure increases, why wouldn’t the density also increase -
especially from this heavier gas? As has been seen, an
increase in density causes cooling, which would then demand
an even greater increase in pressure to offset it, and so on.
If the causative parameters are reversed, and the pressure is
held stable, then density would have to reduce anomalously in
order to reach the required 3°C of warming. In this case;
Pressure: no anomalous change assumed
Density: a decrease of 0.91% due to greenhouse warming
Molar Mass: no anomalous change possible
Calculate for a doubling of CO
2
from the current level of
0.04% (by volume);
 = 101.34
(8.314x 1.2140
28.984)
Calculated temperature after doubling of CO
2
to 0.08%
291.14K
The change in density required, if the pressure remained
stable, would be a fall of 0.91%. This represents an
anomalous change of twenty-three times that which the
percentage change in atmospheric gas alone would logically
indicate. Again, if the pressure fell as well, then the required
fall in density would have to be even greater to compensate.
While still large, the smallest individual anomalous changes
required would be if the pressure rose and simultaneously,
density fell by a similar percentage. Logically, this
combination may be the unlikeliest of these three
possibilities. A possible worked example is provided here;
 = 101.85
(8.314x 1.2197
28.984)
Calculated temperature after doubling of CO
2
to 0.08%
291.16K
This scenario requires an anomalous change of 0.45% to
pressure, combined with an anomalous change of 0.43% to
density. These are eleven times and nine times respectively, the
changes that would reasonably be expected. Evidence of
anomalous changes due to the presence of GHG of this
magnitude are not obvious in the gas data from any of the other
planets, i.e. Venus, Titan. There doesn’t appear to be any
particular class of gases which cause very significant anomalous
changes in any of the gas parameters. This result is not
surprising, since the ideal gas law, in all of its varieties, makes
no distinction between classes of gases based on their radiative
absorption properties. Consistent with this view is that strong
negative feedbacks are evident in the climate system of Earth,
and that there are convincing natural explanations for the recent
period of global warming (see section 3).
2.10. Why the Ideal Gas Law Directly Conflicts with the
Greenhouse Gas Hypothesis
It is known that the ideal gas law does not differentiate
between gases, and so its derivative, the molar mass version
of the ideal gas law cannot either. This fact brings the
derivative into direct conflict with the greenhouse gas
hypothesis and the current, widely-accepted null hypothesis
of climate change. Both of these hypotheses have at their
core, a clear division between gases - those which cause
atmospheric warming and those which do not. It has been
shown that a gas which causes anomalous warming must also
cause anomalous changes to pressure or density or both. Yet
this violates the equivalence of gases which is fundamental to
the ideal gas law. Therefore, either the ideal gas law is
correct, or the greenhouse gas hypothesis of anomalous
warming is correct; both cannot be correct.
A final proof that there can be no anomalous gas parameter
changes due to ‘greenhouse gases’ is that it would be
theoretically possible to change the pressure / density / molar
mass in exactly the same way numerically – by using non-
greenhouse gases to reach the same parameter results – and
the same predicted planetary temperature. Only one
combination of gases is permissible to reach the same
parameter numbers. Therefore, the greenhouse gas hypothesis
and the null hypothesis must be incorrect.
3. Detailed Discussion on the Reasons for
the Late 20
th
Century Warming
What was briefly outlined in previous work [23] and is
being more comprehensively detailed in this work -
essentially for the first time, is the true scientific basis of the
correct null hypothesis for climate change. Arguments have
been aired in the climate literature over the last several
117 Robert Ian Holmes: Thermal Enhancement on Planetary Bodies and the Relevance of the Molar Mass
Version of the Ideal Gas Law to the Null Hypothesis of Climate Change
decades that current global temperatures, and all present
climate change is mostly unnatural, because climate change
is now 98% (according to the IPCC’s reports and its ‘relative
forcings’ chart in AR5) driven by man-made greenhouse
gases, primarily CO
2
[24]. Indeed, it has been claimed that a
distinctively new and wholly man-made geological epoch has
been entered, and that the Holocene now lies in the past [61].
This new epoch has even been named, it is called the
Anthropocene [62].
But there is no solid scientific reason to suppose that
natural climate change does not dominate the present, as it
has always done in the past. For example; no scientific study
has been published in the literature to date, which quantifies
any atmospheric warming and attributes it to increasing man-
made greenhouse gas emissions, primarily CO
2
. It is true that
one paper exists [25] which has quantified a very small
forcing (0.2W/m² in the 2000-2010 period) from increasing
atmospheric CO
2
. It is by no means certain that this has or
will translate into any net warming at all – furthermore, there
is no solid and unchallenged evidence either, that the
measured atmospheric increase in CO
2
in the period 2000-
2010 is all anthropogenic. Yet presently, the widely accepted
‘null’ hypothesis of climate includes significant tropospheric
warming from greenhouse gases. But it is shown in this
work, that if any warming effects might occur in the
troposphere from increases in the greenhouse gas CO
2
, then
they are almost certainly 100% eliminated by negative
feedbacks in the climate system, and so will not manifest in
the troposphere in the form of any net global warming.
Thermodynamics demands that if more CO
2
were to start to
‘create’ an anomalous warming through forcing, then this must
result in atmospheric expansion, because warmer air expands.
But this would increase potential energy at the expense of
kinetic energy - so cooling the air again. The reverse would
also happen; if there were less CO
2
, and this started to cause
cooling, then the atmosphere must contract - so warming the
air again through the conversion of potential to kinetic energy.
Thus, the operation of gas laws coupled with natural
convection are the primary means whereby forcing imbalances
caused by greenhouse gases are eliminated. A second means
of natural negative feedback to a CO
2
forcing in the climate
system arises through cloud production and albedo. More
clouds are known to cause net cooling through a higher albedo
[43]. The effect is not small, a mere 1% change in albedo
being a greater forcing than all the anthropogenic forcing
claimed by the IPCC from 1750 to date.
Opposing the lack of any empirical scientific evidence to
support the claim that man-made GHG are now ‘driving 98%
of global warming and climate change’, is an abundance of
peer-reviewed, published material which supports the
proposed new null hypothesis of natural climate change
continuing right up to the present-day. Most claims about the
anthropogenic nature of climate change focus on the period
since 1950, when almost all anthropogenic emissions of
GHG have occurred. During this period, data from other
planets have also indicated that there has been unusually high
solar activity in recent decades; Mars, [82, 83, 85]; Neptune,
[84]; Pluto, [86, 87].
In particular, focus has generally been on the 1975-2000
warming, since a cooling occurred in the 1950-1975 period
and a slower warming rate was seen in the 2000-2018 period
than in the 1975-2000 period. These periods are well known
and clearly defined in both atmospheric [26] and Oceanic
data [63].
Figure 4. HadCRUT4 data converted to a 50yr trend [26].
Earth Sciences 2018; 7(3): 107-123 118
3.1. The Causes of the 1975-2000 Warming Period
An examination of the HadCRUT4 monthly long-term
global temperature record [88], reveals obvious cyclic peaks
and troughs. The period covered is 1850-2018 and peaks are
seen in 1880, 1940 and 2000. When the data is converted into
a 50yr trend a graph curve is obtained (Figure 4).
The conversion to a 50yr trend clarifies the existence of a
strong ~60yr cycle in the temperature data. (The process of
conversion to this trend transposes the cycle ~15yr towards
the future). This climate cycle is the solar barycentre-related
Yoshimura [27]; it is one among many others seen in the
literature, most of which remain unacknowledged in any of
the IPCC’s reports.
This failure to include relevant science such as this into the
reports strongly biases the conclusions arrived at within them.
Clearly from this chart, the 1910-1940 warming and the 1975-
2000 warming have much in common; yet the IPCC attributes
the first to nature, and the second to man (because there were
few anthropogenic CO
2
emissions during the former, and rapidly
accelerating emissions during the latter). The Yoshimura is in
evidence throughout the climate system, and in proxy records,
on all time-scales [30]. It also seems responsible for the current
Arctic warming, which also occurred in the 1930’s, and was just
as warm then, if not warmer than it is now according the
HadCRUT4 Arctic 70-90N data [28].
A decline of 6% in lower tropospheric tropical cloud cover
(15°N–15°S) occurred 1984 2000 according to the
international satellite cloud climatology project’s data [29].
These years are contained well with the 1975-2000 period of
global warming, and an observed 0.4°C rise in global
temperatures occurred over the same period. Scatter diagrams
[55] of low cloud cover vs global surface air temperatures
indicate that a 1% fall in low clouds equates to a 0.07°C rise
in surface air temperatures - hence this change in cloudiness
accounts for the entire observed rise in global temperatures
during the 1975-2000 period, leaving no room for any effect
from growing greenhouse gases.
3.2. Known Climate Cycles Underpin Current Temperatures
The current period of global warming started ~1690, when
the little ice age bottomed out [33]. Hence, the modern period
of global warming actually started centuries before man-
made emissions could possibly have started to affect the
energy balance. It is seen that there is an underlying increase
in global temperatures in the HadCRUT4 Yoshimura -
influenced record (Figure 4). This underlying 20
th
century
rise, according to recent research on the way temperature
data is processed, may not be due to an energy budget change
at all [89] (Figure 7). This true nature of this underlying rise
is almost certainly natural variability involving the slow
inertial release of stored oceanic heat. The three recent warm
peaks in the 61-year Yoshimura cycle; 1879, 1940 and 2001
are underpinned by other, medium-term climate cycles
including the 248-yr de-Vries, the 1kyr Eddy and the 2.5kyr
Bray (Figure 5) [31, 33]. These medium-term climate cycles
have their origin in solar and planetary orbital interactions,
planetary resonances [30] and barycentre motions [31, 32,
37] the detail of which are too extensive to include here.
Figure 5. The climate cycles which underpin and caused the current period of global warming [33].
How the recent period of global warming (1690-present)
occurred physically, is through a rapidly rising solar forcing
caused by much higher solar activity [34, 35]. The solar
activity in the latter half of the 20
th
century was the highest
for at least 4kyr [36] and perhaps as long as 11kyr [64]. This
initial solar forcing is likely to have been amplified 4-8 times
by feedback mechanisms [37-39] including an albedo-related
one through cosmic rays and low clouds [40, 41]. Forbush
decreases indicate that there is a strong solar-cloud link [42]
though the cosmic ray flux affecting low cloud formation.
Other modes of strong cloud feedbacks are also found in the
climate system. The hemispheric differential in insolation
intensity is known to increase by 15W/m² at the surface over
the period July-January every year due to eccentricity. Yet
this large difference in forcing between the hemispheres
during their respective summers has been measured to be
119 Robert Ian Holmes: Thermal Enhancement on Planetary Bodies and the Relevance of the Molar Mass
Version of the Ideal Gas Law to the Null Hypothesis of Climate Change
virtually eliminated by strong negative cloud feedbacks, by
affecting the relative hemispheric albedo [43]. For a
comparison, the forcing from the increase in CO
2
when
adjusted to span the same time period [25] was measured to
be only 0.01W/m², which is 1,500 times (15/0.01) smaller
than just this one natural change in forcing. Perhaps this
relationship is indicative of just why measuring any influence
on the climate system from increasing atmospheric CO
2
has
proved to be so elusive.
The increase in total solar insolation (TSI) itself is thought
to have been at least several times that reported in the IPCC
reports for the 1750-2000 period according to several papers.
TSI variations must have been at least 3 times larger [44, 45,
46, 64] than is stated in the IPCC reports, otherwise the
severe little ice age cooling centered on 1690 could not have
been possible. Many other published papers [38, 39, 40, 41,
42] show that either the change in TSI since 1750 was much
larger than reported by the IPCC, or there exist strong
amplification mechanisms of that forcing, or (much more
likely) both. The amplification mechanisms also mean that
the 11-yr Schwabe related surface temperature changes are
significant at ~0.2°C [47] and these changes also increase
with height to ~0.8°C in the stratosphere [48].
More evidence supporting the new null hypothesis being
applicable to the late 20
th
century warming comes from a
study of six European city’s thermometer records and other
climate records [49]. The six cities all cover at least the last
230yr, and collectively display no significant overall trend.
However, they do reveal at least five strong climate cycles,
including the 248-yr De-Vries the 80-yr Gleissberg and the
61-yr Yoshimura respectively (Figure 6). The authors
conclude that the cycles themselves explain all global
temperature changes up until the time of publication, (2014)
without any need for a contribution from anthropogenic
greenhouse gases.
Figure 6. Five prominent climate cycles are revealed in the data from six
European cities [49].
Figure 7. Sheltered thermometer stations on land show no warming; 1900-2010 data [89].
Earth Sciences 2018; 7(3): 107-123 120
Further problems for the CO
2
hypothesis arise if cyclic
thermal inertia in the oceans are taken into account, as it has
been in recent work by Lansner & Pedersen [89] (Figure 7).
The current period of global warming started several
centuries ago; when the continuing and slow inertial oceanic
warming from that process is discounted, and only raw
temperature data from sheltered land areas all across the
globe are examined, no warming trend is seen in the period
1900-2010 (Figure 7). Land areas sheltered from ocean
winds currently display temperatures no higher than those
prevailing in the 1920-1950 period.
4. Probable Implications for the Climate
Sensitivity to CO
2
Some reflection upon the simplicity and accuracy of the
planetary results by the use of formula 5, combined with
knowledge of significant other factors such as the common
planetary thermal enhancement / gradient; Venus & Earth
similarities at 1 bar; the supporting material on the 1975-
2000 warming; the reasons why a new null hypothesis of
climate is needed, and the above thought experiment should
enable some probable implications of this work to be
reached. These are that the residual near-surface atmospheric
temperatures on planetary bodies with thick atmospheres are
not mainly determined by the GHE, but instead are very
likely caused by an effect from fluid dynamics, namely; auto-
compression. This leads directly to the conclusion that the
climate sensitivity on Earth to, for example, a doubling of the
atmospheric carbon dioxide concentration has to be not only
operating instantaneously, but also must be extremely low.
On balance, the evidence presented here clearly indicates that
any net temperature change in the lower troposphere, caused
by the addition of 0.04% of CO
2
, cannot be very different to
the addition of a similar quantity of any other gas. In short;
doubling atmospheric CO
2
will not cause a measurable
change in the temperature of the lower troposphere.
The reported figures for equilibrium climate sensitivity to
CO
2
in the literature have already been steadily reducing for
decades, with recent papers pointing to a very low sensitivity
of less than 1°C; [50, 51, 52, 68]. A careful reading of these
papers, (for example the most recent ones) clearly indicates
that the 0.6°C cited, is in fact an absolute maximum. This
present work, if not invalidated by subsequent work, clearly
points to a climate sensitivity so low that it would not be
possible to measure it in the real atmosphere.
To be clear, formulas 5 & 6 when considered in
conjunction with the other material presented here, appears to
rule out any possibility that the assumed 33°C of global
warming from a ‘GHE’ of the type proposed by the IPCC in
their reports can or does exist in the real atmosphere. The
main reason is that the IPCC state in their reports that a
0.04% increase in atmospheric CO
2
, which represents a
doubling from current levels, must result in an average global
lower tropospheric near-surface temperature rise of ~3°C;
(within a range of 1.5°C to 4.5°C) [24, 54] and an even
greater temperature rise at the poles and in the upper
troposphere over the tropics. Atmospheric temperature rises
have not been detected for 40yr over the Antarctic or in the
upper troposphere over the tropics [90]. Despite considerable
new information over recent years, the reported level of
climate sensitivity to a doubling of atmospheric CO
2
, has not
changed significantly from a median 3°C in the regular IPCC
reports since 1990.
Any hypothetical large temperature rises caused by a
doubling of CO
2
must create a large anomalous change in
one or both of two gas parameters (namely, pressure &
density) of which the molar mass version of the ideal gas law
partly consists. There is no supporting scientific evidence for
the existence of these large anomalous changes occurring in
the atmosphere of Earth, or in the atmospheres of other
planetary bodies such as Venus, as a result of a persistently
higher percentages of greenhouse gases.
5. Conclusion
Here is presented a simple and accurate method of
calculating the average near surface atmospheric temperature
on all planetary bodies which possess a surface atmospheric
pressure of over 0.69kPa. This method requires knowledge of
the gas constant and the measurement of only three
atmospheric gas parameters; average near surface
atmospheric pressure, average near surface atmospheric
density and the mean molar mass of the atmosphere.
The formula used is the molar mass version of the ideal
gas law. It is here demonstrated that the information
contained in just these three gas parameters alone is an
extremely accurate predictor of average near-surface
atmospheric temperatures, in all atmospheres >0.69kPa.
Therefore, all information on the effective plus the residual
near-surface atmospheric temperature on planetary bodies
with thick atmospheres, (effective meaning that predicted by
S-B black body law, and residual being the difference
between that and the measured actuality) must be
automatically ‘baked-in’ to these three gas parameters.
A thought experiment involving two planets leads directly
to the conclusion that a small change in any single
atmospheric gas, not only has little effect on atmospheric
temperatures, but has a very similar effect to the same
percentage change in any other atmospheric gas. It is seen
therefore, that as far as this formula goes, no one gas
particularly affects atmospheric temperatures more than any
other gas. Therefore, there can be no significant net
‘greenhouse warming’ caused by ‘greenhouse gases’ on
Earth, or for that matter on any other planetary body. It is
here hypothesised that the residual temperature differences,
and the tropospheric thermal gradients / enhancements
observed on all planetary bodies with thick atmospheres, are
not caused by greenhouse gases. Instead, both are caused by
an effect from thermodynamics, namely a gravity-induced
adiabatic auto-compression of gases, the action of which is to
121 Robert Ian Holmes: Thermal Enhancement on Planetary Bodies and the Relevance of the Molar Mass
Version of the Ideal Gas Law to the Null Hypothesis of Climate Change
enable heat transfer by continuously convecting all gases
within the portions of all atmospheres which are >10kPa. The
mechanism is as follows; rising parcels of air effectively
‘store and hide’ kinetic energy as potential energy; falling
parcels of air effectively ‘convert and reveal’ potential
energy as kinetic energy. The thermodynamics of this process
is described in section 2.3 to 2.5.
If more CO
2
were to start to ‘create’ an anomalous
warming through initiating a forcing, then the laws of
thermodynamics demand that this must result in atmospheric
expansion, because warmer air expands. But this would
increase potential energy at the expense of kinetic energy as
demanded by the above process - so cooling the air again.
The reverse would also happen; if there were less CO
2
, and
this started to cause cooling, then the atmosphere must
contract - so warming the air again through the conversion of
potential to kinetic energy. Thus, the operation of gas laws
coupled with natural convection are the means whereby any
forcing imbalances caused by greenhouse gases are
eliminated.
Further to this, it is suggested that the ‘null’ hypothesis for
climate, as presently understood, is invalid because it
includes a very significant influence from an effect which has
merely been assumed and has not been empirically detected,
quantified or attributed and has not been shown to exist in the
real atmosphere - namely, anomalous tropospheric warming
from so-called ‘greenhouse’ gases. The present ‘null’
hypothesis for climate change needs to be immediately
replaced with one which is based fully on empirical science,
which adheres to the laws of thermodynamics and to the gas
laws, and on atmospheric phenomena which have been
measured, quantified and attributed.
Acknowledgement of non-specific Australian government
support through;
“The Australian Government Research Training Program
Scholarship”.
References
[1] Robinson, T. D., & Catling, D. C. (2014). Common 0.1 [thinsp]
bar tropopause in thick atmospheres set by pressure-dependent
infrared transparency. Nature Geoscience, 7(1), 12-15.
[2] McPherson, M. J. (2012). Subsurface ventilation and
environmental engineering: Springer Science & Business
Media.
[3] Elmegreen, B. G., & Elmegreen, D. M. (1986). Do density
waves trigger star formation? The Astrophysical Journal, 311,
554-562.
[4] Nikolov, N., & Zeller, K. (2017). New insights on the physical
nature of the atmospheric greenhouse effect deduced from an
empirical planetary temperature model. Environment Pollution
and Climate Change, 1(2), 112.
[5] NASA fact sheet data on the planets, (2017). Accessed
11/2/18. https://nssdc.gsfc.nasa.gov /planetary/factsheet/
[6] Seiff, A. (1983). 11. Thermal Structure of the Atmosphere of
Venus. Venus, 215.
[7] Stefan, J. (1879). On the relationship between thermal
radiation and temperature. Bulletin from the sessions of the
Vienna Academy of Sciences (Vienna, 1879), 79, 391-428.
[8] Zasova, L. V., Ignatiev, N., Khatuntsev, I., & Linkin, V.
(2007). Structure of the Venus atmosphere. Planetary and
Space Science, 55(12), 1712-1728.
[9] Wikipedia, Properties of Earth’s atmosphere, (2017).
Accessed 11/2/18.
https://en.wikipedia.org/wiki/Density_of_air
[10] Fulchignoni, M., Ferri, F., Angrilli, F., Ball, A. J., Bar-Nun,
A., Barucci, M. A.,... & Coradini,, M. (2005). In situ
measurements of the physical characteristics of Titan's
environment. Nature, 438(7069), 785-791.
[11] Lindal, G. F., Wood, G., Hotz, H., Sweetnam, D., Eshleman,
V., & Tyler, G. (1983). The atmosphere of Titan: An analysis
of the Voyager 1 radio occultation measurements. Icarus,
53(2), 348-363.
[12] IceCube Wise; Wis/Mad Uni. Accessed 13/11/2017
http://icecube.wisc.edu/pole/weather
[13] Hess, S. L., Henry, R. M., Leovy, C. B., Ryan, J. A., &
Tillman, J. E. (1977). Meteorological results from the surface
of Mars: Viking 1 and 2. Journal of Geophysical Research,
82(28), 4559-4574.
[14] NASA, black body curves Sun and Earth, (2017). Accessed
14/11/2017 https://Earthobser
vatory.nasa.gov/Features/ArcticReflector/Images/black_body_
log_log_rt.gif
[15] Maxwell, J. C. (2012). Theory of heat: Courier Corporation.
[16] Flamm, D. (1997). Four papers by Loschmidt on the state of
thermal equilibrium Pioneering Ideas for the Physical and
Chemical Sciences (pp. 199-202): Springer.
[17] Graeff, R. W. (2007). Viewing The Controversy Loschmidt–
Boltzmann/Maxwell Through Macroscopic Measurements Of
The Temperature Gradients In Vertical Columns Of Water.
Preprint. Additional Results Are on the Web Page.
[18] Arrhenius, S. (1896). XXXI. On the influence of carbonic acid
in the air upon the temperature of the ground. The London,
Edinburgh, and Dublin Philosophical Magazine and Journal
of Science, 41(251), 237-276.
[19] Wood, R. W. (1909). XXIV. Note on the Theory of the
Greenhouse. The London, Edinburgh, and Dublin Philosophical
Magazine and Journal of Science, 17(98), 319-320.
[20] Clough, S. A., Iacono, M. J., & Moncet, J. L. (1992). Line
byline calculations of atmospheric fluxes and cooling rates:
Application to water vapor. Journal of Geophysical Research:
Atmospheres, 97(D14), 15761-15785.
[21] Khilyuk, L. (2003). Global warming: are we confusing cause
and effect? Energy Sources, 25(4), 357-370.
[22] Feldman, D. R., Collins, W. D., Gero, P. J., Torn, M. S.,
Mlawer, E. J., & Shippert, T. R. (2015). Observational
determination of surface radiative forcing by CO
2
from 2000
to 2010. Nature, 519(7543), 339-343.
[23] Holmes, R. I. (2017c). Molar Mass Version of the Ideal Gas
Law Points to a Very Low Climate Sensitivity. Earth Sciences,
6(6), 157.
Earth Sciences 2018; 7(3): 107-123 122
[24] Team, C. W., Pachauri, R., & Meyer, L. (2014). IPCC, 2014:
Climate Change 2014: Synthesis Report. Contribution of
Working Groups I. II and III to the Fifth Assessment Report of
the Intergovernmental Panel on Climate Change. IPCC,
Geneva, Switzerland, 151.
[25] Feldman, D. R., Collins, W. D., Gero, P. J., Torn, M. S.,
Mlawer, E. J., & Shippert, T. R. (2015). Observational
determination of surface radiative forcing by CO
2
from 2000
to 2010. Nature, 519(7543), 339-343.
[26] Climate4you 50yr trend, (2018). Access 11/2/18
http://www.climate4you.com/images/HadCRUT4%2050yr%2
0AnnualitresendSinceDecember1899.gif
[27] Yoshimura, H. (1979). The solar-cycle period-amplitude
relation as evidence of hysteresis of the solar-cycle nonlinear
magnetic oscillation and the long-term/55 year/cyclic
modulation. The Astrophysical Journal, 227, 1047-1058.
[28] http://www.climate4you.com/images/70-
90N%20MonthlyAnomaly%20Since1920.gif Accessed
11/2/2018.
[29] https://isccp.giss.nasa.gov/products/onlineData.html Accessed
11/2/2018.
[30] Semi, P. (2009). Orbital resonance and Solar cycles. arXiv
preprint arXiv:0903.5009.
[31] Velasco, V., Mendoza, B., & Valdes-Galicia, J. (2008). The
120-yrs solar cycle of the cosmogenic isotopes. Paper
presented at the International Cosmic Ray Conference.
[32] Scafetta, N. (2014). The complex planetary synchronization
structure of the solar system. arXiv preprint arXiv:1405.0193.
[33] Steinhilber, F., Abreu, J. A., Beer, J., Brunner, I., Christl, M.,
Fischer, H.,... McCracken, K. G. (2012). 9,400 years of
cosmic radiation and solar activity from ice cores and tree
rings. Proceedings of the National Academy of Sciences,
109(16), 5967-5971.
[34] Lüdecke, H.-J. (2011). Long-term instrumental and
reconstructed temperature records contradict anthropogenic
global warming. Energy & Environment, 22(6), 723-745.
[35] Solanki, S. K., Usoskin, I. G., Kromer, B., Schüssler, M., &
Beer, J. (2004). Unusual activity of the Sun during recent
decades compared to the previous 11,000 years. Nature,
431(7012), 1084-1087.
[36] Usoskin, I. G., Solanki, S., & Kovaltsov, G. (2007). Grand
minima and maxima of solar activity: new observational
constraints. Astronomy & Astrophysics, 471(1), 301-309.
[37] Shaviv, N. J. (2008). Using the oceans as a calorimeter to
quantify the solar radiative forcing. Journal of Geophysical
Research: Space Physics, 113(A11).
[38] Svensmark, H. (2007b). Cosmoclimatology: a new theory
emerges. Astronomy & Geophysics, 48(1), 1.18-11.24.
[39] Shaviv, N. J. (2005). On climate response to changes in the
cosmic ray flux and radiative budget. Journal of Geophysical
Research: Space Physics, 110(A8).
[40] Svensmark, H., Enghoff, M. B., & Pedersen, J. O. P. (2013).
Response of cloud condensation nuclei (> 50 nm) to changes
in ion-nucleation. Physics Letters A, 377(37), 2343-2347.
[41] Yu, F., & Luo, G. (2014). Effect of solar variations on particle
formation and cloud condensation nuclei. Environmental
Research Letters, 9(4), 045004.
[42] Svensmark, H., Bondo, T., & Svensmark, J. (2009). Cosmic
ray decreases affect atmospheric aerosols and clouds.
Geophysical Research Letters, 36(15).
[43] Stephens, G. L., O'Brien, D., Webster, P. J., Pilewski, P., Kato,
S., & Li, J.-l. (2015). The albedo of Earth. Reviews of
Geophysics, 53(1), 141-163.
[44] Foukal, P., Fröhlich, C., Spruit, H., & Wigley, T. M. L. (2006).
Variations in solar luminosity and their effect on the Earth's
climate. Nature, 443(7108), 161-166.
[45] Douglass, D. H., & Clader, B. D. (2002). Climate sensitivity
of the Earth to solar irradiance. Geophysical Research Letters,
29(16).
[46] Pulkkinen, T., Nevanlinna, H., Pulkkinen, P., & Lockwood, M.
(2001). The Sun–Earth connection in time scales from years to
decades and centuries. Space Science Reviews, 95(1-2), 625-637.
[47] Camp, C. D., & Tung, K. K. (2007). Surface warming by the
solar cycle as revealed by the composite mean difference
projection. Geophysical Research Letters, 34(14).
[48] White, W. B. (2006). Response of tropical global ocean
temperature to the Sun's quasidecadal UV radiative forcing
of the stratosphere. Journal of Geophysical Research: Oceans,
111(C9).
[49] Lüdecke, H.-J., Hempelmann, A., & Weiss, C. (2013). Multi-
periodic climate dynamics: spectral analysis of long-term
instrumental and proxy temperature records. Climate of the
Past, 9(1), 447.
[50] Harde, H. (2014). Advanced Two-Layer Climate Model for
the Assessment of Global Warming by CO
2
.
[51] Cederlöf, M. (2014). Using seasonal variations to estimate
Earth's response to radiative forcing.
[52] Abbot, J., & Marohasy, J. (2017). The application of machine
learning for evaluating anthropogenic versus natural climate
change. GeoResJ, 14, 36-46.
[53] http://ossfoundation.us/projects/environment/global-
warming/atmospheric-composition Accessed 11/2/2018.
[54] Allen, M. R., Barros, V. R., Broome, J., Cramer, W., Christ,
R., Church, J. A.,... Dubash, N. K. (2014). IPCC Fifth
Assessment Synthesis Report-Climate Change 2014 Synthesis
Report.
[55] http://www.climate4you.com/images/TotalCloudCoverVersus
GlobalSurfaceAirTemperature.gif accessed 13/2/2018.
[56] Graeff, R. W. (2002, November). Measuring the temperature
distribution in gas columns. In AIP Conference
Proceedings(Vol. 643, No. 1, pp. 225-230). AIP.
[57] Levy, G. S. (2016, June). Loschmidt’s Temperature Gradient
Paradox—A Quantum Mechanical Resolution. In 97th Annual
Meeting of the American Association for the Advancement of
Science, Limits to the Second Law of Thermodynamics,
University of San Diego, San Diego, California (pp. 14-17).
[58] Moroz, V., Ekonomov, A., Moshkin, B., Revercomb, H.,
Sromovsky, L., Schofield, J., Tomasko, M. G. (1985). Solar
and thermal radiation in the Venus atmosphere. Advances in
Space Research, 5(11), 197-232.
123 Robert Ian Holmes: Thermal Enhancement on Planetary Bodies and the Relevance of the Molar Mass
Version of the Ideal Gas Law to the Null Hypothesis of Climate Change
[59] Zasova, L., Ignatiev, N., Khatuntsev, I., & Linkin, V. (2007).
Structure of the Venus atmosphere. Planetary and Space
Science, 55(12), 1712-1728.
[60] Lindal, G. F., Hotz, H. B., Sweetnam, D. N., Shippony, Z.,
Brenkle, J. P., Hartsell, G. V., & Michael, W. H. (1979).
Viking radio occultation measurements of the atmosphere and
topography of Mars: Data acquired during 1 Martian year of
tracking. Journal of Geophysical Research: Solid Earth,
84(B14), 8443-8456.
[61] Lewis, S. L., & Maslin, M. A. (2015). Defining the
anthropocene. Nature, 519(7542), 171.
[62] Zalasiewicz, J., Williams, M., Steffen, W. & Crutzen, P
(2010). The new world of the Anthropocene.
[63] Gouretski, V., Kennedy, J., Boyer, T., & Köhl, A. (2012).
Consistent nearsurface ocean warming since 1900 in two
largely independent observing networks. Geophysical
Research Letters, 39(19).
[64] Yndestad, H., & Solheim, J. E. (2017). The influence of solar
system oscillation on the variability of the total solar
irradiance. New Astronomy, 51, 135-152.
[65] Harde, H. (2017). Scrutinizing the carbon cycle and CO 2
residence time in the atmosphere. Global and Planetary
Change, 152, 19-26.
[66] Lacis, A. A., Schmidt, G. A., Rind, D., & Ruedy, R. A.
(2010). Atmospheric CO2: Principal control knob governing
Earth’s temperature. Science, 330(6002), 356-359.
[67] Cook, J., Oreskes, N., Doran, P. T., Anderegg, W. R.,
Verheggen, B., Maibach, E. W.,... Green, S. A. (2016).
Consensus on consensus: a synthesis of consensus estimates
on human-caused global warming. Environmental Research
Letters, 11(4), 048002.
[68] Lightfoot, H. D., & Mamer, O. A. (2014). Calculation of
atmospheric radiative forcing (Warming Effect) of carbon
dioxide at any concentration. Energy & Environment, 25(8),
1439-1454.
[69] Lepori, L., Bussolino, G., Matteoli, E., & Spanedda, A. On the
increase of fossil CO2 in the atmosphere.
[70] Segalstad, T. V. (1998). Carbon cycle modelling and the
residence time of natural and anthropogenic atmospheric CO
2
.
BATE, R.(Ed., 1998): Global Warming, 184-219.
[71] Quirk, T. (2009). Sources and sinks of carbon dioxide. Energy
& Environment, 20(1), 105-121.
[72] Pauli, W. (1988). Exclusion principle, Lorentz group and
reflection of space-time and charge Wolfgang Pauli (pp. 459-
479): Springer.
[73] Hollingsworth, J., Young, R., Schubert, G., Covey, C., &
Grossman, A. (2007). A simplephysics global circulation
model for Venus: Sensitivity assessments of atmospheric
superrotation. Geophysical Research Letters, 34(5).
[74] Trenberth, K. E., Fasullo, J. T., & Kiehl, J. (2009). Earth's
global energy budget. Bulletin of the American Meteorological
Society, 90(3), 311-323.
[75] Landis, G., Dyson, R., McGuire, M., Oleson, S., Schmidt, G.,
Grantier, J.,... Fincannon, J. (2011). Human Telerobotic
Exploration of Venus: A Flexible Path Design Study. Paper
presented at the 49th AIAA Aerospace Sciences Meeting
including the New Horizons Forum and Aerospace
Exposition.
[76] Svedhem, H., Titov, D. V., Taylor, F. W., & Witasse, O.
(2007). Venus as a more Earth-like planet. Nature, 450(7170),
629-632.
[77] Jelbring, H. (2003). The “Greenhouse Effect” as a Function of
Atmospheric Mass. Energy & Environment, 14(2), 351-356.
[78] Pätzold, M., Häusler, B., Bird, M. K., Tellmann, S., Mattei, R.,
Asmar, S. W.,... & Tyler, G. L. (2007). The structure of Venus’
middle atmosphere and ionosphere. Nature, 450(7170), 657.
[79] Crisp, D. (2007). Greenhouse Effect and Radiative Balance on
Earth and Venus. Presentation to the Venus Exploration
Assessment Group (VEXAG).
[80] Pierrehumbert, R. T. (2011, November). Infrared radiation and
planetary temperature. In AIP Conference Proceedings (Vol.
1401, No. 1, pp. 232-244). AIP.
[81] Porco, C. C., Baker, E., Barbara, J., Beurle, K., Brahic, A.,
Burns, J. A.,... & Denk, T. (2005). Imaging of Titan from the
Cassini spacecraft. Nature, 434(7030), 159.
[82] Fenton, L., Geissler, P., & Haberle, R. (2006). Global
warming on Mars. Paper presented at the AGU Fall Meeting
Abstracts.
[83] Pollack, J. B., Kasting, J. F., Richardson, S. M., & Poliakoff,
K. (1987). The case for a wet, warm climate on early Mars.
Icarus, 71(2), 203-224.
[84] Sromovsky, L., Fry, P., Limaye, S., & Baines, K. (2003). The
nature of Neptune’s increasing brightness: Evidence for a
seasonal response. Icarus, 163(1), 256-261
[85] Ravilious, K. (2007). Mars melt hints at solar, not human,
cause for warming, scientist says. National Geographic News.
http://news. nationalgeogr.../55741367. html.
[86] Pasachoff, J. M., Souza, S. P., Babcock, B. A., Ticehurst, D.
R., Elliot, J., Person, M., Tholen, D. J. (2005). The structure of
Pluto's atmosphere from the 2002 August 21 stellar
occultation. The Astronomical Journal, 129(3), 1718.
[87] Elliot, J. L., Person, M., Gulbis, A., Souza, S., Adams, E.,
Babcock, B.,... Pasachoff, J. (2007). Changes in Pluto’s
atmosphere: 1988-2006. The Astronomical Journal, 134(1), 1.
[88] Climate Research Unit data;
https://crudata.uea.ac.uk/cru/data/temperature/HadCRUT4-
gl.dat Accessed 10/3/18.
[89] Lansner, F., & Pedersen, J. O. P. (2017). Temperature trends
with reduced impact of ocean air temperature.
[90] University of Alabama at Huntsville data from the Trios-N
satellite 1979-2018.
https://www.nsstc.uah.edu/data/msu/t2lt/uahncdc.lt Accessed
10/3/2018
... It is known that all planetary bodies with thicker atmospheres naturally set up a rising thermal gradient in that part of the atmosphere, which is higher than a pressure of 0.1 bar, until that bodies' surface is reached [1]. Previous works [2][3][4]14] have indicated that there may be a relationship between total solar irradiance (TSI), atmospheric pressures and planetary atmospheric temperature on bodies which possess thick atmospheres. It is shown here that this relationship appears to exist across all three terrestrial-type bodies which possess thick atmospheres. ...
... It is known from the Stefan-Boltzmann law that the radiating temperature of an isolated planetary body in space, (one which possesses no atmosphere), varies with the fourthroot of the power incident upon it [30]. And given that previous works [1][2][3][4] have detailed that a principle factor in determining atmospheric temperatures on planetary bodies with thick atmospheres is atmospheric pressure, logic dictates that these may be combined -initially at a standard for pressure -for example, 1 bar. ...
... This result is consistent with previous research [3] with regards to atmospheric temperatures and their relationship to the molar mass version of the ideal gas law, in that this work also points to a climate sensitivity to CO₂ -or to any other socalled 'greenhouse' gas -which is close to or at zero. It is more confirmation that the main determinants of atmospheric temperatures in the regions of terrestrial planetary atmospheres which are >0.1 bar, is overwhelmingly the result of two factors; solar insolation and atmospheric pressure. ...
Article
Full-text available
It has been discovered that there appears to exist a close relationship between relative differences in total solar irradiance and the atmospheric temperature, at a pressure of 1 bar, on all three terrestrial-type bodies which possess thick atmospheres. The apparent relationship is through the quaternary root of total solar irradiance at 1 bar, and applies to the planetary bodies Venus, Earth and Titan. The relationship is so close that the average surface atmospheric temperature of Earth can be easily calculated to within 1 Kelvin (0.5%) of the correct figure by the knowledge of only two numbers, neither of which are related to the Earth's atmosphere. These are; the atmospheric temperature in the Venusian atmosphere at 1 bar, and the top-of-atmosphere solar insolation of the two planets. A similar relationship in atmospheric temperatures is found to exist, through insolation differences alone, between the atmospheric temperatures at 1 bar of the planetary bodies Titan and Earth, and Venus and Titan. This relationship exists despite the widely varying atmospheric greenhouse gas content, and the widely varying albedos of the three planetary bodies. This result is consistent with previous research with regards to atmospheric temperatures and their relationship to the molar mass version of the ideal gas law, in that this work also points to a climate sensitivity to CO2-or to any other 'greenhouse' gas-which is close to or at zero. It is more confirmation that the main determinants of atmospheric temperatures in the regions of terrestrial planetary atmospheres which are >0.1 bar, is overwhelmingly the result of two factors; solar insolation and atmospheric pressure. There appears to be no measurable, or what may be better termed 'anomalous' warming input from a class of gases which have up until the present, been incorrectly labelled as 'greenhouse' gases.
... It is known that all planetary bodies with thicker atmospheres naturally set up a rising thermal gradient in that part of the atmosphere, which is higher than a pressure of 0.1 bar, until that bodies' surface is reached [1]. Previous works [2,3,4,14] have indicated that there may be a relationship between total solar irradiance (TSI), atmospheric pressures and planetary atmospheric temperature on bodies which possess thick atmospheres. It is shown here that this relationship appears to exist across all three terrestrial-type bodies which possess thick atmospheres. ...
... It is known from the Stefan-Boltzmann law that the radiating temperature of an isolated planetary body in space, (one which possesses no atmosphere), varies with the fourth-root of the power incident upon it [30]. And given that previous works [1,2,3,4] have detailed that a principle factor in determining atmospheric temperatures on planetary bodies with thick atmospheres is atmospheric pressure, logic dictates that these may be combined -initially at a standard for pressure -for example, 1 bar. ...
... The data shows that the 'greenhouse gas' concentration varies widely from the low 2.7% and 2.5% [5,9,10,11] for Titan and Earth respectively, to the very high 96.5% for Venus; the implication must be that there cannot be any special warming effect from the so-called 'greenhouse' gases. This result adds and contributes to considerable other evidence [2,3,4,14,17,18,24,25,26,27,28,29], that there is no sign of any 'greenhouse effect' from 'greenhouse' gases on any of these three bodies. ...
Preprint
Full-text available
It has been discovered that there appears to exist a close relationship between relative differences in total solar irradiance and the temperature of all three terrestrial-type bodies which possess thick atmospheres, at a pressure of 1 bar. The apparent relationship is through the quaternary root of total solar irradiance at 1 bar, and applies to the planetary bodies Venus, Earth and Titan.
... Solid particles are efficient radiators of thermal energy and so our modelling study suggests a causal link between the bright sulphur veil albedo of Venus and its atmospheric thickness. 7. The requirement for a tropopause at a temperature that allows for the formation of solid particles of a planet's main condensing volatile, which for Venus is sulphuric acid, indicates that planetary Bond Albedo is an interlinked consequence of atmospheric mass and thermal profile [18]. Consequently, Bond Albedo is an emergent property and not a cause of a planet's atmospheric thermal profile, because the freezing point requirement to form solid particles of a condensing volatile specifically determines the planet's TOA thermal emission temperature. ...
Preprint
Full-text available
The terrestrial planet Venus is classified by astronomers as an inferior planet because it is located closer to the Sun than the Earth. Venus orbits the Sun at a mean distance of 108.21 Million Km and receives an average annual solar irradiance of 2601.3 W/m 2 , which is 1.911 times that of the Earth. A set of linked forward and inverse climate modelling studies were undertaken to determine whether a process of atmospheric energy retention and recycling could be established by a mechanism of energy partition between the solid illuminated surface and an overlying fully transparent, non-greenhouse gas atmosphere. Further, that this atmospheric process could then be used to account for the observed discrepancy between the average annual solar insolation flux and the surface tropospheric average annual temperature for Venus. Using a geometric climate model with a globular shape that preserves the key fundamental property of an illuminated globe, namely the presence on its surface of the dual environments of both a lit and an unlit hemisphere; we established that the internal energy flux within our climate model is constrained by a process of energy partition at the surface interface between the illuminated ground and the overlying air. The dual environment model we have designed permits the exploration and verification of the fundamental role that the atmospheric processes of thermal conduction and convection have in establishing and maintaining surface thermal enhancement within the troposphere of this terrestrial planet. We believe that the duality of energy partition ratio between the lit and unlit hemispheres applied to the model, fully accounts for the extreme atmospheric “greenhouse effect” of the planet Venus. We show that it is the meteorological process of air mass movement and energy recycling through the mechanism of convection and atmospheric advection, associated with the latitudinal hemisphere encompassing Hadley Cell that accounts for the planet’s observed enhanced atmospheric surface warming. Using our model, we explore the form, nature and geological timing of the climatic transition that turned Venus from a paleo water world into a high-temperature, high-pressure carbon dioxide world.
... When gas is compressed, it does negative work, and the temperature rises because some of the potential energy is converted into enthalpy, H through H = PxV + U, where U is the specific internal energy (kinetic energy) of the system in Joules. A molecule or ion can absorb some of the energy up to a certain point (Holmes, 2017(Holmes, , 2018Thousendfrolly, 2017). Electrons can reach a higher state, and chemical reactions can occur, thereby establishing a new equilibrium. ...
Preprint
Full-text available
Economic- and ecological plan in (/outside) European countries: Income=Spending+Annuity[interest and redemption national debt]*. Solution for the financial-, energy-, climate-, trade- and food crisis.
... Solid particles are efficient radiators of thermal energy and so our modelling study suggests a causal link between the bright sulphur veil albedo of Venus and its atmospheric thickness. 7. The requirement for a tropopause at a temperature that allows for the formation of solid particles of a planet's main condensing volatile, which for Venus is sulphuric acid, indicates that planetary Bond Albedo is an interlinked consequence of atmospheric mass and thermal profile [18]. Consequently, Bond Albedo is an emergent property and not a cause of a planet's atmospheric thermal profile, because the freezing point requirement to form solid particles of a condensing volatile specifically determines the planet's TOA thermal emission temperature. ...
Article
Full-text available
The terrestrial planet Venus is classified by astronomers as an inferior planet because it is located closer to the Sun than the Earth. Venus orbits the Sun at a mean distance of 108.21 Million Km and receives an average annual solar irradiance of 2601.3 W/m 2 , which is 1.911 times that of the Earth. A set of linked forward and inverse climate modelling studies were undertaken to determine whether a process of atmospheric energy retention and recycling could be established by a mechanism of energy partition between the solid illuminated surface and an overlying fully transparent, non-greenhouse gas atmosphere. Further, that this atmospheric process could then be used to account for the observed discrepancy between the average annual solar insolation flux and the surface tropospheric average annual temperature for Venus. Using a geometric climate model with a globular shape that preserves the key fundamental property of an illuminated globe, namely the presence on its surface of the dual environments of both a lit and an unlit hemisphere; we established that the internal energy flux within our climate model is constrained by a process of energy partition at the surface interface between the illuminated ground and the overlying air. The dual environment model we have designed permits the exploration and verification of the fundamental role that the atmospheric processes of thermal conduction and convection have in establishing and maintaining surface thermal enhancement within the troposphere of this terrestrial planet. We believe that the duality of energy partition ratio between the lit and unlit hemispheres applied to the model, fully accounts for the extreme atmospheric "greenhouse effect" of the planet Venus. We show that it is the meteorological process of air mass movement and energy recycling through the mechanism of convection and atmospheric advection, associated with the latitudinal hemisphere encompassing Hadley Cell that accounts for the planet's observed enhanced atmospheric surface warming. Using our model, we explore the form, nature and geological timing of the climatic transition that turned Venus from a paleo water world into a high-temperature, high-pressure carbon dioxide world.
... Few papers could be found in the worldwide literature which singularly address how to reduce the remaining VAM emissions specifically for environmental reasons in an operating colliery, even though these emissions represent most of a typical colliery's 'greenhouse' gas emissions (Zhongqing et al., 2010) (Su et al., 2005). However, later publications show that anomalous warming from 'greenhouse' gases may not exist in the real atmosphere (Holmes, 2018). ...
Preprint
Full-text available
A twelve-month VAM mitigation trial was conducted at a Hunter Valley, NSW colliery in 2012-2013 using a method involving six measures, which aimed to prevent some methane from entering the mine airstream and becoming VAM. The six measures combined, reduced emissions by a quantified total of 95,398 t/CO₂-e below that which was projected for the next twelve-month period at the low cost of A$1.08 t/CO₂-e. Safety was also considered in the trial, and each measure was individually found to have increased safety somewhat at the mine. Presented here for the first time, is work that was conducted at another colliery in Queensland in 2016, aimed at validating or invalidating whether this mitigation method is transferable to other collieries. One of the more successful of the six measures used during the NSW trial, involved the pressure balancing a sealed panel; this measure was chosen to be the subject of this work.
Preprint
Full-text available
Considering the period from 1750 to 1850, the CO2 concentration was approximately 275 ppm in 1750 and approximately 284.7 ppm in 1850. However, the temperature in 1750 is estimated to be approximately 0.5-1.1 °C higher than that in 1850. Thus, the temperature decreased from 1750 to 1850, as the CO2 level monotonic increased from approximately 275 ppm in 1750 to approximately 284.7 ppm in 1850. Therefore, there is no correlation between the CO2 concentration and the temperature. Natural cycles such as the day and night cycle (amplitude 2,11 - 6,59 ° C), the summer and winter cycle (amplitude 5,81 ° C) , the 11-year solar cycle (amplitude 0,1 ° C), the 65-year (Atlantic Multidecadal Oscillation (AMO)/Pacific Decadal Oscillation (PDO)) cycle (amplitude 0.15 ° C), the 200-year de Vries / Suess cycle (amplitude 0.4 ° C) and the 1000-year Eddy cycle (amplitude 0.4 ° C) are responsible for the temperature on Earth. With the ideal gas law, the temperature on planetary bodies with a thick atmosphere above 0.1 bar can be calculated. The Equilibrium climate sensitivity (ECS) of CO2 from 284.7 ppm to 569.4 ppm at 101.325 kPa on the surface is approximately -0.000808 K and I proved this in extensive novel calculations in an excel file. CO2 can only absorb a certain amount of energy up to a certain point, after that energy is released, energy is not trapped as suggested in the greenhouse gas theory. Our earth is an open system and does not act like a blanket or greenhouse which is a closed system. I proved this with extensive calculations at 50 kPa (≈5,960 meter height) and 10kPa (≈19,540 meter height) in an excel file.
Chapter
Änderungen des Klimas, z. T. mit deutlich höheren Temperaturen als heute, durchziehen die gesamte Erdgeschichte. Da fragt man sich, ob nicht auch natürliche anstatt menschlicher Ursachen der Erwärmung in Frage kommen. Wir gehen dieser Frage nach und diskutieren den Zusammenhang zwischen CO2, Wasserdampf und Klima in der Erdgeschichte und heute, die Rolle der Ozeane und der Sonne, um mit der Klimasensitivität einen zentralen Begriff der Klimadiskussion aufzugreifen.
Preprint
Full-text available
Considering the period from 1750 to 1850, the CO2 concentration was approximately 275 ppm in 1750 and approximately 284.7 ppm in 1850. However, the temperature in 1750 is estimated to be approximately 0.5-1.1 °C higher than that in 1850. Thus, the temperature decreased from 1750 to 1850, as the CO2 level monotonic increased from approximately 275 ppm in 1750 to approximately 284.7 ppm in 1850. Therefore, there is no correlation between the CO2 concentration and the temperature. Natural cycles such as the day and night cycle (amplitude 2,11 - 6,59 ° C), the summer and winter cycle (amplitude 5,81 ° C) , the 11-year solar cycle (amplitude 0,1 ° C), the 65-year (Atlantic Multidecadal Oscillation (AMO)/Pacific Decadal Oscillation (PDO)) cycle (amplitude 0.15 ° C), the 200-year de Vries / Suess cycle (amplitude 0.4 ° C) and the 1000-year Eddy cycle (amplitude 0.4 ° C) are responsible for the temperature on Earth. With the ideal gas law, the temperature on planetary bodies with a thick atmosphere above 0.1 bar can be calculated. The Equilibrium climate sensitivity (ECS) of CO2 from 284.7 ppm to 569.4 ppm at 101.325 kPa on the surface is approximately -0.000808 K. Do you want to support research by Het Landschap, you can do this by transferring 25 EURO annually in a standing order to account number NL73 ABNA 0825 5566 43. of W.P.A.H. de Vries with the notice of Het Landschap.
Article
Full-text available
It has always been complicated mathematically, to calculate the average near surface atmospheric temperature on planetary bodies with a thick atmosphere. Usually, the Stefan Boltzmann (S-B) black body law is used to provide the effective temperature, then debate arises about the size or relevance of additional factors, including the ‘greenhouse effect’. Presented here is a simple and reliable method of accurately calculating the average near surface atmospheric temperature on planetary bodies which possess a surface atmospheric pressure of over 10kPa. This method requires a gas constant and the knowledge of only three gas parameters; the average near-surface atmospheric pressure, the average near surface atmospheric density and the average mean molar mass of the near-surface atmosphere. The formula used is the molar version of the ideal gas law. It is here demonstrated that the information contained in just these three gas parameters alone is an extremely accurate predictor of atmospheric temperatures on planets with atmospheres >10kPa. This indicates that all information on the effective plus the residual near-surface atmospheric temperature on planetary bodies with thick atmospheres, is automatically ‘baked-in’ to the three mentioned gas parameters. Given this, it is shown that no one gas has an anomalous effect on atmospheric temperatures that is significantly more than any other gas. In short; there can be no 33°C ‘greenhouse effect’ on Earth, or any significant ‘greenhouse effect’ on any other planetary body with an atmosphere of >10kPa. Instead, it is a postulate of this hypothesis that the residual temperature difference of 33°C between the S-B effective temperature and the measured near-surface temperature is actually caused by adiabatic auto-compression.
Article
Full-text available
A recent study has revealed that the Earth’s natural atmospheric greenhouse effect is around 90 K or about 2.7 times stronger than assumed for the past 40 years. A thermal enhancement of such a magnitude cannot be explained with the observed amount of outgoing infrared long-wave radiation absorbed by the atmosphere (i.e. ≈ 158 W m-2), thus requiring a re-examination of the underlying Greenhouse theory. We present here a new investigation into the physical nature of the atmospheric thermal effect using a novel empirical approach toward predicting the Global Mean Annual near-surface equilibrium Temperature (GMAT) of rocky planets with diverse atmospheres. Our method utilizes Dimensional Analysis (DA) applied to a vetted set of observed data from six celestial bodies representing a broad range of physical environments in our Solar System, i.e. Venus, Earth, the Moon, Mars, Titan (a moon of Saturn), and Triton (a moon of Neptune). Twelve relationships (models) suggested by DA are explored via non-linear regression analyses that involve dimensionless products comprised of solar irradiance, greenhouse-gas partial pressure/density and total atmospheric pressure/density as forcing variables, and two temperature ratios as dependent variables. One non-linear regression model is found to statistically outperform the rest by a wide margin. Our analysis revealed that GMATs of rocky planets with tangible atmospheres and a negligible geothermal surface heating can accurately be predicted over a broad range of conditions using only two forcing variables: top-of-the-atmosphere solar irradiance and total surface atmospheric pressure. The hereto discovered interplanetary pressure-temperature relationship is shown to be statistically robust while describing a smooth physical continuum without climatic tipping points. This continuum fully explains the recently discovered 90 K thermal effect of Earth’s atmosphere. The new model displays characteristics of an emergent macro-level thermodynamic relationship heretofore unbeknown to science that has important theoretical implications. A key entailment from the model is that the atmospheric ‘greenhouse effect’ currently viewed as a radiative phenomenon is in fact an adiabatic (pressure-induced) thermal enhancement analogous to compression heating and independent of atmospheric composition. Consequently, the global down-welling long-wave ux presently assumed to drive Earth’s surface warming appears to be a product of the air temperature set by solar heating and atmospheric pressure. In other words, the so-called ‘greenhouse back radiation’ is globally a result of the atmospheric thermal effect rather than a cause for it. Our empirical model has also fundamental implications for the role of oceans, water vapour, and planetary albedo in global climate. Since produced by a rigorous attempt to describe planetary temperatures in the context of a cosmic continuum using an objective analysis of vetted observations from across the Solar System, these ndings call for a paradigm shift in our understanding of the atmospheric ‘greenhouse effect’ as a fundamental property of climate.
Article
Full-text available
Total solar irradiance (TSI) is the primary quantity of energy that is provided to the Earth. The properties of the TSI variability are critical for understanding the cause of the irradiation variability and its expected influence on climate variations. A deterministic property of TSI variability can provide information about future irradiation variability and expected long-term climate variation, whereas a non-deterministic variability can only explain the past. This study of solar variability is based on an analysis of two TSI data series, one since 1700 A.D. and one since 1000 A.D.; a sunspot data series since 1610 A.D.; and a solar orbit data series from 1000 A.D. The study is based on a wavelet spectrum analysis. First, the TSI data series are transformed into a wavelet spectrum. Then, the wavelet spectrum is transformed into an autocorrelation spectrum to identify stationary, subharmonic and coincidence periods in the TSI variability. The results indicate that the TSI and sunspot data series have periodic cycles that are correlated with the oscillations of the solar position relative to the barycenter of the solar system, which is controlled by gravity force variations from the large planets Jupiter, Saturn, Uranus and Neptune. A possible explanation for solar activity variations is forced oscillations between the large planets and the solar dynamo. We find that a stationary component of the solar variability is controlled by the 12-year Jupiter period and the 84-year Uranus period with subharmonics. For TSI and sunspot variations, we find stationary periods related to the 84-year Uranus period. Deterministic models based on the stationary periods confirm the results through a close relation to known long solar minima since 1000 A.D. and suggest a modern maximum period from 1940 to 2015. The model computes a new Dalton-type sunspot minimum from approximately 2025 to 2050 and a new Dalton-type period TSI minimum from approximately 2040 to 2065.
Article
Full-text available
The consensus that humans are causing recent global warming is shared by 90%–100% of publishing climate scientists according to six independent studies by co-authors of this paper. Those results are consistent with the 97% consensus reported by Cook et al (Environ. Res. Lett. 8 024024) based on 11 944 abstracts of research papers, of which 4014 took a position on the cause of recent global warming. A survey of authors of those papers (N = 2412 papers) also supported a 97% consensus. Tol (2016 Environ. Res. Lett. 11 048001) comes to a different conclusion using results from surveys of non-experts such as economic geologists and a self-selected group of those who reject the consensus. We demonstrate that this outcome is not unexpected because the level of consensus correlates with expertise in climate science. At one point, Tol also reduces the apparent consensus by assuming that abstracts that do not explicitly state the cause of global warming ('no position') represent non-endorsement, an approach that if applied elsewhere would reject consensus on well-established theories such as plate tectonics. We examine the available studies and conclude that the finding of 97% consensus in published climate research is robust and consistent with other surveys of climate scientists and peer-reviewed studies.
Article
Temperature data 1900–2010 from meteorological stations across the world have been analyzed and it has been found that all land areas generally have two different valid temperature trends. Coastal stations and hill stations facing ocean winds are normally more warm-trended than the valley stations that are sheltered from dominant oceans winds. Thus, we found that in any area with variation in the topography, we can divide the stations into the more warm trended ocean air-affected stations, and the more cold-trended ocean air-sheltered stations. We find that the distinction between ocean air-affected and ocean air-sheltered stations can be used to identify the influence of the oceans on land surface. We can then use this knowledge as a tool to better study climate variability on the land surface without the moderating effects of the ocean. We find a lack of warming in the ocean air sheltered temperature data – with less impact of ocean temperature trends – after 1950. The lack of warming in the ocean air sheltered temperature trends after 1950 should be considered when evaluating the climatic effects of changes in the Earth’s atmospheric trace amounts of greenhouse gasses as well as variations in solar conditions.
Article
This book has been written as a reference and text for engineers, researchers, teachers and students who have an interest in the planning and control of the environment in underground openings. While directed primarily to underground mining operations, the design procedures are also applicable to other complex developments of subsurface space such as nuclear waste repositories, commercial accommodation or vehicular networks. The book will, therefore, be useful for mining, civil, mechanical, and heating, ventilating and air-conditioning engineers involved in such enterprises. The chapters on airborne pollutants highlight means of measurement and control as well as physiological reaction. These topics will be of particular interest to industrial hygienists and students of industrial medicine. One of the first technical applications of digital computers in the world's mining industries was for ventilation network analysis. This occurred during the early 1960s. However, it was not until low cost but powerful personal computers proliferated in engineering offices during the 1980s that the full impact of the computer revolution was realized in the day-to-day work of most mine ventilation engineers. This book reflects the changes in approach and design procedures that have been brought about by that revolution. While the book is organized into six parts, it encompasses three broad areas.
Article
Time-series profiles derived from temperature proxies such as tree rings can provide information about past climate. Signal analysis was undertaken of six such datasets, and the resulting component sine waves used as input to an artificial neural network (ANN), a form of machine learning. By optimizing spectral features of the component sine waves, such as periodicity, amplitude and phase, the original temperature profiles were approximately simulated for the late Holocene period to 1830 AD. The ANN models were then used to generate projections of temperatures through the 20th century. The largest deviation between the ANN projections and measured temperatures for six geographically distinct regions was approximately 0.2°C, and from this an Equilibrium Climate Sensitivity (ECS) of approximately 0.6°C was estimated. This is considerably less than estimates from the General Circulation Models (GCMs) used by the Intergovernmental Panel on Climate Change (IPCC), and similar to estimates from spectroscopic methods.
Article
Climate scientists presume that the carbon cycle has come out of balance due to the increasing anthropogenic emissions from fossil fuel combustion and land use change. This is made responsible for the rapidly increasing atmospheric CO2 concentrations over recent years, and it is estimated that the removal of the additional emissions from the atmosphere will take a few hundred thousand years. Since this goes along with an increasing greenhouse effect and a further global warming, a better understanding of the carbon cycle is of great importance for all future climate change predictions. We have critically scrutinized this cycle and present an alternative concept, for which the uptake of CO2 by natural sinks scales proportional with the CO2 concentration. In addition, we consider temperature dependent natural emission and absorption rates, by which the paleoclimatic CO2 variations and the actual CO2 growth rate can well be explained. The anthropogenic contribution to the actual CO2 concentration is found to be 4.3%, its fraction to the CO2 increase over the Industrial Era is 15% and the average residence time 4 years.
Chapter
The 70th anniversary of Niels Bohr’s birthday reminds me of a long and still continuing common pilgrimage since the year 1922, in which so many stations are involved. Without pretension of completeness I mention here only some of them in their relation to the particular subject of this paper which, I hope, he will permit me to dedicate to him on the occasion of this celebration.