- Home
- Research Overview
- Contact
- Summary Page 1
- Research Page 1
- Research Page 2
- Research Page 3
- Research Page 4
- Research Page 5
- Research Page 6
- Research Page 7
- Research Page 8
- Research Page 9
- Research Page 10
- Research Page 11
- Research Page 12
- Research Page 13
- Research Page 14
- Research Page 15
- Research Page 16
- Research Page 17
- Research Page 18
- Research page 19
- Research Page 20
- Research Page 21
- Climate Humor Page 1
- Climate Humor Page 2
REJECT AR6
Ventura Photonics Climate Post 006.1a Feb. 20 2022
Roy Clark
The Sixth IPCC Climate Assessment Report (AR6) should be rejected outright because it is based on the use of fraudulent climate models. This fraud comes from the underlying assumption of radiative forcing in an equilibrium climate used to construct the models. Such models are fraudulent by definition, before a single line of computer code is written. Climate science has now degenerated past scientific dogma into the quasi-religious ‘Imperial Cult of the Global Warming Apocalypse’. Scientific reason has been replaced by blind advocacy. There is no ‘climate crisis’. Eisenhower’s warning about the corruption of science by government funding has come true. The entire multi-trillion dollar Ponzi or pyramid scheme built on these fraudulent modeling results needs to be shut down and those responsible should face the legal consequences of their activities.
SUMMARY
The recently published draft of the latest UN Intergovernmental Panel on Climate Change (IPCC) Report, Climate Change 2021: The Physical Science Basis, [IPCC, 2021] the contribution of Working Group 1 to the Sixth IPCC Climate Assessment (AR6) should be rejected outright because the report is based on the results from fraudulent climate models. This fraud comes from the underlying assumption of radiative forcing in an equilibrium average climate used to construct the climate models. These models are fraudulent by definition before any computer code is even written. AR6 is a continuation of the climate modeling fraud that started with the invalid assumptions that were made when the first computer climate models were developed in the 1960s. All of the equilibrium climate model results used by the IPCC since it was established in 1988 are fraudulent.
There are at least three separate parts to this fraud. First there are the invalid climate equilibrium and related assumptions that originated in the nineteenth century. These led to melodramatic prophecies of the global warming apocalypse and became such a good source of research funding that the scientific process of hypothesis and discovery collapsed. Second, there was institutional fraud related to ‘mission creep’ within various government agencies. For example, NASA was established to put a man on the moon. There was no provision to shut it down after that mission was accomplished. Climate modeling provided alternative employment for some of the NASA ‘scientists’ with nothing else to do. The climate fraud was firmly established at NASA by 1981. Third, there was a deliberate decision by various outside interests, including environmentalists and politicians to exploit the climate apocalypse to further their own causes. There was no single person or event that created the climate fraud. There was a gradual transition from the invalid hypothesis of an equilibrium average climate to the massive multi-trillion dollar pyramid or Ponzi scheme that we have today.
The peer review process in climate science has collapsed and been replaced by blatant cronyism. The climate modelers have retreated inside a cocoon of lies where they discuss the pseudoscience of radiative forcings, feedbacks and climate sensitivities to a CO2 ‘doubling’. This is just climate theology. How does a doubling of the atmospheric CO2 concentration change the number of angels that may dance on the head of a climate pin? Here they use their ‘advanced’ climate models to create the sacred spaghetti plots of global warming. This is GIGO: garbage in, gospel out. The model ‘predictions’ are fed directly to government agencies and the IPCC with minimal outside scrutiny. There is no climate science involved. Irrational belief in computer models has replaced the Laws of Physics. The Imperial Cult of the Global Warming Apocalypse has claimed the Divine Right to save the world from a non-existent problem. There is no ‘climate crisis’. Eisenhower’s warning about the corruption of science by government funding has come true. The entire multi-trillion dollar Ponzi or pyramid scheme built on these fraudulent modeling results needs to be shut down and those responsible should face the legal consequences of their activities.
TECHNICAL SUMMARY
The fundamental scientific error in the climate models is the climate equilibrium assumption. This has been used to oversimplify the energy transfer processes that determine the surface temperature. It creates the illusion that the surface temperature is determined by the LWIR flux. In reality, a change in surface temperature has to be determined from the change in enthalpy or heat content of the surface thermal reservoir divided by the heat capacity. Any small change in LWIR flux from a so called ‘radiative forcing’ has to be added to the rest of the flux terms in an interactive, time dependent thermal engineering calculation of the surface temperature. There are four main flux terms, the solar heating, the net LWIR emission, the moist convection (evapotranspiration) and the subsurface thermal transport. Furthermore, the LWIR flux cannot be separated and analyzed independently of the other flux terms. It is fully coupled to the moist convection.
There is an abundance of evidence that clearly demonstrates that the equilibrium climate assumption is invalid. However, this has been ignored by the climate modeling community. Fourier described time delays or phase sifts between the peak solar flux and the subsurface seasonal temperature response in 1824. Such phase shifts can be found in the seasonal and daily temperature changes in both land and ocean temperatures. They are definitive evidence of non-equilibrium thermal response. These phase shifts alone are sufficient to invalidate the equilibrium climate models. In addition, there are also quasi-periodic ocean oscillations that provide a ‘noise floor’ for climate temperature variations.
The idea that changes in the atmospheric concentration of CO2 could cause the earth to cycle through an Ice Age was first proposed by Tyndall in the 1860s. This was finally disproved in 1976 when the 100,000 year Ice Age cycle was linked to changes in the earth’s orbital eccentricity through the analysis of deep drilled ocean sediment cores. If changes in atmospheric CO2 concentration do not cycle the earth through an Ice Age then there is no reason to expect that changes in CO2 concentration from fossil fuel combustion should have any effect on climate.
Simple energy balance calculations show that the long term planetary average LWIR flux returned to space should be near 240 W m-2. Satellite radiometer measurements give a value of 240±100 W m-2. However, this is simply a cooling flux that does not define an emission temperature. The spectral distribution is not that of a blackbody near 255 K. This means that there can be no ‘equilibrium greenhouse effect temperature’ of 33 K. All of the ‘discussion’ over a ‘greenhouse effect temperature’ is nothing more than climate theology.
The underlying assumption used in equilibrium climate modeling is that there exists an ‘equilibrium average climate’ with an exact flux balance at the top of the atmosphere (TOA) between the absorbed solar flux and the LWIR flux emitted to space. An increase in atmospheric CO2 concentration reduces the LWIR flux within the CO2 emission bands. This is called a ‘radiative forcing'. It is then assumed that this is a perturbation to the climate equilibrium state. The surface temperature is then supposed to ‘adjust’ to a new higher temperature to restore the flux balance at TOA. It is also assumed that the surface temperature response is linear and that there is an amplification of the warming by a ‘water vapor feedback’. A doubling of the CO2 concentration from a ‘preindustrial level’ of 280 parts per million (ppm) to 560 ppm produces a ‘radiative forcing’ of 3.7 W m-2. A pseudoscientific ‘climate sensitivity’ to CO2 is then used to create an increase in ‘equilibrium surface temperature’ of 3.7 ±1.9 C.
In reality, a change in flux produces a change in the rate of heating or cooling, not a change in temperature. The ‘radiative forcing’ at TOA is produced by absorption at lower levels in the atmosphere mainly by the P and R braches of the main CO2 band near 640 and 700 cm-1. The maximum increase in the rate of heating for a ‘CO2 doubling’ is near 0.08 K per day at an altitude near 2 km. This is coupled to the normal convective motion in the troposphere. At an average lapse rate of 6.5 K km-1 this corresponds to a decrease in altitude of about 12 meters. This is equivalent to riding an elevator down about 4 floors. The small amount of heat produced at each level in the troposphere is dissipated by a combination of broadband emission, mainly by the water bands and a slight increase in the gravitational potential energy. It does not couple to the surface. There is no change to the ‘energy balance’ of the earth, the absorbed LWR flux is just rearranged and emitted at different wavelengths.
Although it is not explicitly discussed by the IPCC as part of radiative forcing, there is also an increase in the downward LWIR flux to the surface that is similar in magnitude to the reduction in LWIR flux at TOA. The radiation field in the atmosphere consists of many thousands of overlapping lines. Each line corresponds to a transition between two vibration-rotation states of a so called ‘greenhouse gas’. The dominant greenhouse gas molecules are water vapor and CO2. In the lower troposphere the lines are pressure broadened. Within the main absorption-emission bands, the lines merge to form a quasi-continuous blackbody source. Because of these line broadening effects, almost all of the downward LIWR flux to the surface originates from within the first 2 km layer above the surface and approximately half of the LWIR flux originates from within the first 100 m layer. This means that the LWIR emission to space is decoupled from the downward LWIR flux to the surface.
The troposphere functions as an open cycle heat engine that divides naturally into two independent thermal reservoirs. The lower reservoir extends from the surface to an altitude near 2 km. The upper thermal reservoir extends from 2 km to the tropopause. The upper reservoir is the cold reservoir of the heat engine. As the warm moist air rises from the surface by convection, the air expands and cools. The some of the internal molecular heat energy is converted to gravitational potential energy. As the air cools by LWIR emission at higher altitudes in the troposphere, it sinks and gravitational potential energy is converted back to heat. If the air is moist, water vapor condenses above the saturation level. Clouds form and latent heat is released. For dry air, the lapse rate or temperature profile for convective ascent is -9.8 K km-1. For moist air the magnitude of the lapse is reduced because of the latent heat release. The US standard atmosphere uses an average moist lapse rate of -6.5 K km-1. In addition, air can be compressed and heated by a downward flow. This occurs in high pressure systems and with downslope winds. Near surface air temperatures may change by 10 C over a short time period from hours to a few days.
The greenhouse effect is simply the LWIR exchange energy at the surface. The downward LWIR flux from the lower troposphere ‘blocks’ most of the upward LWIR flux the surface. There is an exchange of photons but much reduced heat transfer. In order to dissipate the absorbed solar flux, the surface must warm up so that the excess heat is dissipated by moist convection. In addition, some of the heat may be transported and stored below the surface. The land and ocean surfaces have different thermal properties and have to be analyzed separately.
Over land, almost all of the absorbed solar heat is dissipated within the same diurnal cycle. As the surface warms during the day, the excess heat is removed by moist convection. Some of the heat is conducted below the surface, stored and returned to the surface later in the day. In the evening, the surface cools and the convection essentially stops as the surface and air temperatures equalize. The surface then cools more slowly over night by net LWIR emission. The equalization or convection transition temperature is reset each day by the local weather system passing through. The day to day changes in the convection transition temperature are sufficiently large that any change in surface temperature from an in increase in the downward LWIR flux from CO2 is too small to measure. There can be no CO2 induced temperature ‘signal’ in the weather station data used to create the average surface temperature anomaly.
Over the oceans, the water surface is almost transparent to the solar flux. Approximately 90% of the solar flux is absorbed within the first 10 m layer of the ocean. The diurnal temperature rise is small and the bulk ocean temperature increases until the excess solar heat is dissipated by wind driven evaporation. The cooler surface water sinks and is replaced by warmer water from below. This allows the evaporation to continue at night. Within the ±30° latitude bands the sensitivity of the evaporation to the wind speed is approximately 15 W m-2/m s-1. The latent heat flux increases by 15 W m-2 for an increase in wind speed of 1 m s-1. The penetration depth of the LWIR flux into the ocean surface is 100 micron or less. Here the LWIR flux is fully coupled to the wind driven cooling flux. The two should not be separated and analyzed separately. There can be no measurable ocean warming from CO2. The magnitude and variability in the wind driven latent heat flux are too large. The average increase in atmospheric CO2 concentration is approximately 2.4 ppm per year. This produces an increase in downward LWIR flux to the surface of 0.034 W m-2. This is equivalent to a change in wind speed of 2.3 mm per year.
1.0 INTRODUCTION
Over the last 200 years, the atmospheric concentration of CO2 has increased by approximately 130 ppm from 280 to 410 ppm as shown in Figure 1a [Keeling, 2021]. This has been attributed to increases in fossil fuel combustion since the start of the industrial revolution [Meinshausen et al, 2011]. Countries such as China and India are still building new coal fired electrical power generation plants, so the atmospheric CO2 concentration will continue to increase [BP, 2020]. The LWIR flux in the atmosphere can be calculated using high resolution radiative transfer algorithms and HITRAN or a similar spectroscopic database [HITRAN, 2020, Wijngaarden and Happer, 2020]. The results of such calculations from Harde [2017], Hansen [2005] and Clark [2013] are shown in Figure 1b. This shows the change in total LWIR flux. Most of this occurs within the P and R branches of the v2 CO2 band near 640 and 700 cm-1. There is also a small contribution from the overtone bands near 950 and 1050 cm-1. There is good agreement between the different calculations. As the CO2 concentration increases, there is a slight increase in the downward LWIR flux reaching the surface and a small decrease in the LWIR flux emitted at TOA. For the observed increase of 130 ppm, the change in flux is approximately 2 W m-2. For a doubling of the CO2 concentration from ‘preindustrial’ levels of 280 ppm to 560 ppm, the decrease at TOA is given as 3.7 W m-2 [IPCC 2013] or 3.9 W m-2 [IPCC, 2021] and for a doubling from 360 to 720 ppm, the decrease is near 5 W m-2. The calculated change in LWIR flux may vary slightly depending on the surface temperature and humidity values selected. Currently, the average increase in atmospheric CO2 concentration is near 2.4 ppm per year and the increase in downward flux to the surface is approximately 0.034 W m-2 per year.
Figure 1: a) The increase in atmospheric CO2 concentration from 1800 and b) calculated changes in atmospheric LWIR flux produced by an increase in atmospheric CO2 concentration from 35 to 760 ppm.
The issue therefore is not the value of the change in atmospheric LWIR flux produced by an increase in atmospheric CO2 concentration, but the effect of this change on the earth’s climate, starting with the calculation of the increase in surface temperature. Here there is complete disagreement between the conventional thermal engineering approach and the equilibrium radiative forcing techniques used in the climate models [Poyet, 2020, Gerlich and Tscheuschner, 2009].
In the real world, the surface temperature is determined by the interaction of four main time dependent heat flux terms with the surface thermal reservoir. These are the absorbed solar flux, the net LWIR emission, the evapotranspiration (moist convection) and the subsurface thermal transport. (Rainfall and freeze/thaw effects are not included here). A change in temperature is determined by dividing the change in heat content or enthalpy by the local heat capacity of the thermal reservoir. In order to determine the effect of an increase in atmospheric CO2 concentration on the surface temperature, the change in enthalpy has to be determined after a solar thermal cycle with increased downward LWIR flux from CO2. There are two parts to this analysis. First, the change in temperature from the small increase in downward LWIR flux from CO2 has to be evaluated. Second, the effect of other processes such as changes in local weather conditions have to be evaluated to determine if the CO2 induced changes are even measurable. In signal processing terms, this is the determination of the signal to noise ratio. The results of such calculations show that the increase in surface temperature produced by the increase in downward LWIR flux from CO2 is too small to measure.
The climate models start from the invalid concept of an equilibrium average climate. It is assumed that there is an exact long term planetary energy balance between the average absorbed solar flux and the average outgoing LWIR radiation (OLR) returned to space. An increase in the atmospheric CO2 concentration produces a slight reduction the LWIR flux at TOA within the CO2 emission band. This is considered to be a perturbation to the equilibrium state. The re-emission of the absorbed LWIR flux at other wavelengths and the coupling to the convection are ignored. The climate is then presumed to ‘adjust’ so that there is an increase in ‘equilibrium surface temperature’ that restores the flux balance at TOA [Knutti, and Hegerl, 2008]. The change in flux at TOA is called a radiative forcing. It is further assumed that there is a linear relationship between the ‘radiative forcing’ and the surface temperature response.
[Harde, 2017, IPCC, 2013 Chap. 8]. Climate sensitivity is defined in terms of the temperature change produced by doubling of the CO2 concentration from a ‘preindustrial’ level of 280 ppm to 560 ppm. It is assumed a-priori that all of the recent changes in temperature record must be attributable to the increase in ‘radiative forcing’. Effects such as natural oscillations in ocean surface temperatures are ignored.
2.0 ENERGY TRANSFER IN A NON-EQUILIBRIUM CLIMATE SYSTEM
In order to understand how the earth’s climate really works it necessary to introduce four concepts that are not part of a ‘climate equilibrium state’.
1) The so called ‘greenhouse effect’ has to be defined as the time dependent surface LWIR exchange energy.
2) There are significant phase shifts or time delays in the climate system between the peak solar flux and the surface temperature response.
3) The penetration depth of the LWIR flux into the oceans is approximately 100 micron. Here it is fully coupled to the wind driven evaporation or latent heat flux.
4) The land surface temperature is reset each day by the convection transition temperature.
These concepts will now considered in more detail.
2.1 The Greenhouse Effect
The earth’s surface is warmer than it should be based on rather simple energy conservation arguments. The average solar flux at the top of the atmosphere (TOA) is near 1368 W m-2. The exact value depends on satellite radiometer calibration [Wilson, 2014]. The albedo or reflectivity is near 0.3. The geometry is that of a sphere illuminated by a circular beam of nearly collimated solar radiation. The sphere to disk surface area ratio is four. This means that the average LWIR flux returned to space should be near 1368*(1-0.3)/4 ≈ 240 W m-2. IPCC AR6 gives 239 (237 to 242) W m-2. If the earth were a uniform blackbody emitter, this would correspond to an emission temperature near 255 K. Simple inspection of IR satellite images of the earth, such as the CERES image shown in Figure 2 indicates that the LWIR flux is approximately 240 ±100 W m-2 [CERES, 2011]. The various heating and cooling rates within the climate system interact to keep the surface temperature within the relatively narrow range needed to sustain the development of life on earth. These rates are always changing on diurnal, seasonal and longer time scales. There is no equilibrium.
Figure 2: CERES IR image of the earth, March 8, 2011
Furthermore, the spectral distribution of the LWIR flux emitted by the earth at the top of the atmosphere is not that of a blackbody. There is no ‘shell’ of gas surrounding the earth with a temperature near 255 K. The outgoing LWIR radiation (OLR) at TOA is simply the cumulative emission from many different layers of the atmosphere. The emission from each layer is modified by the absorption and emission of the layers above. These temperatures are also changing because of convective (vertical) motion in the troposphere. The spectral distribution of the OLR under ‘clear sky’ conditions is illustrated in Figure 3. The LWIR flux consists of a mix of atmospheric emission, mainly from the H2O and CO2 bands and surface emission through the LWIR transmission window. Some of the surface emission is absorbed by stratospheric ozone. The main spectral features are labelled. Blackbody emission curves at selected temperatures are also shown. The 255 K blackbody emission curve is shown as the black dotted line. It has no relationship to the TOA flux shown by the orange line. The LWIR flux at TOA is just a cooling flux that should not be used to define an ‘effective emission temperature’. This means that there can be no greenhouse effect temperature near 33 K [Taylor, 2006].
Figure 3: The LWIR flux emitted at TOA (orange line). The blackbody emission curves at selected temperatures are also shown. This TOA flux cannot be described as LWIR blackbody emission near 255 K. MODTRAN calculation, 288 K surface temperature, 80% RH, 400 ppm CO2, 100 to 1500 cm-1 spectral range, 2 cm-1 spectral resolution, mid latitude summer [MODTRAN, 2020].
It is also well known that the OLR at ‘clear sky’ increases linearly with surface temperature as shown in Figure 4 [Koll and Cronin, 2018]. There are two contributions to the OLR flux with different responses to the surface temperature. Within the LWIR transmission window in the 800 to 1200 cm-1 region, the surface emission is only partially absorbed and some is emitted to space. Here, the clear sky OLR increases linearly with surface temperature. Within the main H2O and CO2 absorption bands, the OLR emission does not change significantly with surface temperature. The water band emission profile shifts to higher altitude as the surface temperature increases. More heat is stored as gravitational potential energy. The LWIR absorption and emission process continues with increasing altitude until the molecular linewidths narrow sufficiently to allow the transition to a free photon flux to space. For H2O, this transition occurs near a temperature of 253 K (-20 C). For CO2 the free photon transition occurs at a lower temperature near 220 K (-53 C). Most of the CO2 band emission occurs in the stratosphere. This is shown below in Figure 12.
Figure 4: The increase in clear sky OLR flux with surface temperature
The observed linear increase in OLR flux in response to an increase in surface temperature does not mean that a change in OLR flux can couple back to the surface and change the surface temperature. This is a violation of the Second Law of Thermodynamics. Instead, the greenhouse effect is simply the result of the LWIR exchange energy at the surface. The downward LWIR flux from the lower troposphere ‘balances’ most the upward LWIR flux from the surface. Photons are exchanged without a net transfer of heat. When the surface and air temperatures are similar, the net LWIR flux limited to the LWIR transmission window, mainly in the 800 to 1200 cm-1 spectral region. The net LWIR flux increases with decreasing humidity and decreases with increasing cloud cover. This is illustrated in Figure 5. In order to dissipate the excess absorbed solar insolation, the surface warms up until the heat is removed by moist convection (evapotranspiration). This is a non-equilibrium process in which some of the absorbed solar heat is stored in the surface thermal reservoir and then released as the rates of heating and cooling change during the diurnal and seasonal cycles. In addition, some of the solar heat is stored below the ocean surface and can be transported over long distances by ocean currents. The energy transfer processes at the land-air and ocean–air are different, so the two have to be analyzed separately.
Figure 5: a) The surface exchange energy and b) MODTRAN calculation, using the same conditions as Figure 3.
The troposphere functions as an open cycle heat engine that transports part of the absorbed solar heat from the surface to the middle and upper troposphere by moist convection. From here it is radiated to back space, mainly by the water bands. Some of the surface heat is stored as gravitational potential energy in the troposphere. Convection is a mass transport process that is coupled to both the gravitational potential and the angular momentum or rotation of the earth. These interactions result in the formation of the Hadley, Ferrell and polar cell convective structure, the trade winds and the ocean gyre circulation. The ‘greenhouse effect’ is an integral part of the energy transfer processes that determine the earth’s weather patterns. The LWIR flux in the troposphere is part of the tropospheric heat engine and should not be separated and analyzed independently from the mass transport. The greenhouse effect is produced by the time dependent surface LWIR exchange energy on a rotating water planet.
2.2 The Time Delay or Phase Shift between the Peak Solar Flux and the Temperature Response
In order to warm or cool the surface thermal reservoir, heat must flow into or out of the reservoir. This introduces a time delay or phase shift between the peak solar flux and the temperature response. This type of phase shift also occurs on different time and energy scales in other energy storage devices. These include capacitors and inductors in electronic AC circuits and photon storage in passive optical resonators (cavity ringdown). This is not new science. The seasonal subsurface phase shift was described by Fourier in 1824 and 1827 [Fourier, 1824, 1827]. The most obvious phase shift is the seasonal delay between the peak solar insolation at summer solstice and the peak summer temperatures. At mid latitudes, this is usually between 4 and 8 weeks. Such phase shifts have been recorded by the weather stations for well over 100 years. Figure 6 shows the phase shifts for selected weather stations in S. California. These are the 30 year daily average MSAT Tmin and Tmax 1981-2010 climate data for Los Angeles, LA Airport (LAX), Redlands, Riverside, San Bernardino, Blythe, Indio and Mecca [WRCC, 2020]. The values of the phase shifts in days after summer solstice are shown in Figure 7. The observation of such phase shifts provides clear evidence of non-equilibrium thermal storage. This is sufficient to show that there is no equilibrium climate state that can be perturbed by an increase in the atmospheric CO2 concentration.
Figure 6: 30 year daily average MSAT Tmin and Tmax 1981-2010 climate data for selected weather stations in S. California. The seasonal phase shifts past summer solstice are clearly visible.
Figure 7: Seasonal phase shifts (days after summer solstice) for the weather station climate data shown in Figure 6.
Such phase shifts are also found in ocean temperature data. In fact the seasonal phase shifts are first generated by ocean solar heating. They are coupled to the weather station measurements through the diurnal convective transition temperature as the weather systems move over land. Figure 8 shows the monthly 2.5 m to 200 m depth ocean temperature profiles for 2018 for the N. Atlantic Ocean at 30° N, 20° W derived from Argo float data [Argo 2020]. The data are for a 5° x 1° (latitude x longitude) strip. The minimum winter surface temperature of 19 C is reached in March. Solar heating then produces a stable stratified thermal layer structure with a surface temperature peak of 24 C in August. The subsurface thermal layer structure then collapses as the wind driven evaporative cooling in winter exceeds the solar heating. The time delay or phase shift between the peak solar flux and the peak surface temperature response is approximately 8 weeks. These phase shifts increase with depth. At 60 m the phase shift is 16 weeks.
Figure 8: The 2.5 m to 200 m depth ocean temperature profiles for 2018 for the N. Atlantic Ocean at 30° N, 20° W derived from Argo float data [Argo 2020]. The data are for a 5° x 1° (latitude x longitude) strip. The seasonal phase shifts are indicated.
In addition to the seasonal phase shifts, there are also diurnal phase shifts in both the land and ocean temperatures that may reach 2 hours or more. However, such phase shifts were not usually recorded as part of the historical weather station record.
2.3 The Coupling of the LWIR Flux to the Oceans
The penetration depth of the LWIR flux into the ocean surface is approximately 100 micron (0.004 in). This is illustrated in Figure 9 [Hale and Querry, 1973]. Here it is fully coupled to the wind driven surface evaporation or latent heat flux. The sensitivity of the ocean latent heat flux to the wind speed may be estimated from the long term zonal data provided by Yu et al [2008]. This is shown in Figure 10. Over the ±30° latitude bands, the sensitivity is at least 15 W m-2/m s-1. As shown above in Figure 1b, the increase in downward LWIR flux to the surface produced by the observed 120 ppm increase in atmospheric CO2 concentration is approximately 2 W m-2. Within the ±30° latitude bands, this is dissipated by an increase in wind speed near 13 cm s-1. The average increase in CO2 concentration at present is near 2.4 ppm. This corresponds to an annual increase of 0.034 W m-2 in the downward LWIR flux to the surface. This is dissipated by an increase in wind speed near 2 millimeters per second.
Figure 9: Penetration depth of LWIR radiation into the oceans, a) 3300 to 50 cm-1 and b) 1200 to 200 cm-1. The locations of the main CO2 band and the overtones are indicated.
Figure 10: The sensitivity of the ocean latent heat flux to the wind speed.
2.4 The Convection Transition Temperature
Over land, almost all of the absorbed solar flux is dissipated within the same diurnal cycle. The surface temperature is reset each evening by the convection transition temperature at which the surface and air temperatures start to equalize. This depends on the local weather conditions. The transition temperature may be understood by examining the diurnal flux terms and surface and surface air temperature changes for a dry surface under full summer sun illumination shown schematically in Figure 11. These are based on measurements recorded at the ‘Grasslands’ site located in Limestone Canyon, near Irvine, S. California [Clark, 2013]. The flux is positive if there is a heat flow into the surface. The peak surface temperature of 50 C and peak surface air temperature of 25 C are reached approximately 2 hours after solar noon. This time delay or phase shift is clear evidence of a non-equilibrium thermal response. The transition temperature is reached in the late evening and the surface then cools more slowly by net LWIR emission. The net LWIR flux decreases to 50 W m-2 at night and increases to a maximum value of 218 W m-2 during the day. This is insufficient to dissipate the absorbed solar heat, so the surface temperature adjusts until the excess heat is removed by convection. Heat is also conducted below the surface during the first part of the day after sunrise. This heat flow reverses during the afternoon and the absorbed heat is returned to the surface. In this example, approximately 60% of the absorbed solar flux is dissipated as convection.
Figure 11: Diurnal variation in a) flux terms and b) surface and air temperatures for a dry surface under full summer sun conditions. Schematic based on ‘Grasslands’ data from UC Irvine [Clark, 2013].
3.0 RADIATIVE FORCING
As shown above in Figure 1b, the observed increase of 120 ppm in the atmospheric CO2 concentration has produced a decrease in the LWIR flux emitted to space within the CO2 emission bands of approximately 2 W m-2. There has also been a similar increase in the downward LWIR flux to the surface. The radiative forcing argument used by the IPCC assumes that the decrease in LWIR flux at TOA perturbs the earth’s radiation balance. The climate system is then supposed to respond by adjusting to a new ‘equilibrium state’ with an increase in surface temperature that restores the LWIR flux balance at TOA [IPCC, 2021, Chapter 7, IPCC, 2013, Chapter 8, Knutti and Hegerl, 2008]. In order to match the temperature record, the initial temperature rise produced by a ‘radiative forcing’ then has to be amplified by a ‘water vapor feedback’. This is pseudoscientific nonsense. In order understand the real energy transfer processes involved, it is necessary to consider the upward and downward LWIR flux terms separately, including the coupling to the mass transport.
3.1 The Decrease in LWIR Flux at TOA
The decrease in LWIR flux at TOA is produced by an increase in absorption by CO2 at lower levels in the atmosphere. Since there is no thermal equilibrium, the change in absorption has to be converted to changes in the local rate of heating. This involves the calculation of the absorbed LWIR flux at each level in the atmosphere and dividing by the local heat capacity. In addition, it is necessary to consider molecular line broadening effects in the lower troposphere and the coupling of the LWIR flux to the mass transport (convection) in the troposphere. Most of the initial absorption occurs in the P and R branches of the v2 CO2 band near 640 and 700 cm-1. There is also some absorption by the CO2 overtone bands near 950 and 1050 cm-1. The slight warming produced by these absorptions is then dissipated by a combination of wideband LWIR emission across all of the atmospheric emission bands and coupling to the convection. Some of the thermal energy is converted to gravitational potential energy followed by LWIR emission at a later time.
The total and band averaged cooling rate profiles for the tropical model atmosphere are shown in Figure 12. In the lower and middle troposphere the total cooling rate is near 2 K per day [Feldman et al, 2008, Lacis and Oindas, 1991]. The changes in atmospheric heating rates for a ‘doubling’ of the CO2 concentration, in this case from 287 to 574 ppm at mid latitude are shown in Figure 13 [Iacono et al, 2008]. In the troposphere, the maximum increase in heating rate is less than 0.1 K per day at an altitude near 2 km. In the stratosphere, there is a maximum increase in cooling rate near 50 km of 3 K per day.
Figure 12: Total and band-averaged IR cooling rate profiles for the Tropical Model Atmosphere on a log-pressure scale [data from Feldman et al, 2008].
Figure 13: a) Tropospheric heating and b) stratospheric cooling rates produced by a CO2 ‘doubling’ from 287 to 574 ppm at mid latitude [data from Iacono et al. 2008].
The LWIR absorption and emission in the atmosphere consists of thousands of overlapping lines, each involving transitions between specific molecular rotation-vibration states [Clark, 2013, Wijngaarden and Happer, 2020]. The dominant molecular species are H2O and CO2. Through the troposphere and most of the stratosphere the molecular linewidth is determined by pressure broadening. This is the result of molecular collisions. Near the surface, within the main absorption bands the individual lines merge to form a quasi-continuous blackbody. The absorption and the molecular linewidths decrease as the pressure, temperature and the species concentration decrease with altitude. The upward and downward LWIR fluxes through the atmosphere are not equivalent. Some of the upward flux from the wings of the broader lines below can pass through the gaps between the lines above. Conversely the downward flux from the narrower lines above is absorbed by the wider lines below. The change in linewidth with altitude is illustrated in Figure 14a. Almost all of the downward flux reaching the surface from the main absorption bands is emitted from within the first 2 km layer above the surface. Approximately half of this flux is emitted by the first 100 m layer above the surface. It is this downward flux that provides the photons for the surface exchange energy. The cumulative downward flux from H2O and CO2 vs. altitude is shown in Figure 14b. Four cases are plotted for surface temperatures of 272 and 300 K each with relative humidities of 20 and 70%. The downward flux near the surface increases with temperature and humidity. Even for the lowest flux case, 272 K and 20% RH, 95% of the surface flux originates from within the first 2 km layer. The downward flux to the surface is decoupled from the LWIR emission to space. The troposphere divides naturally into two thermal reservoirs. The lower reservoir extends from the surface to 2 km and the upper reservoir extends from 2 km to the tropopause.
Figure 14: a) transition from absorption-emission to free photon flux as the linewidth decreases with altitude, changes in linewidth for a single H2O line near 231 cm-1. b) Cumulative fraction of the downward flux at the surface vs. altitude for surface temperatures of 272 and 300 K, each with 20 and 70% RH. Almost all of the downward flux reaching the surface originates from within the first 2 km layer.
The troposphere functions as an open cycle heat engine that transports heat from the surface to higher altitudes in the troposphere by moist convection. From here it is radiated back to space, mainly by the water bands. The local temperature profile of the troposphere is set by the local lapse rate, which depends on the surface temperature, the relative humidity and the convection. The stratosphere is heated mainly by the absorption of UV solar flux by ozone and cools by LWIR emission from CO2 and ozone. The local solar heating changes on a daily and a seasonal time scale. The downward flux from the LWIR emission in the stratosphere and upper troposphere is absorbed in the lower troposphere and does not reach the surface. The local temperature of an air parcel in the troposphere depends on the local flux balance. Within the plane parallel atmosphere approximation there are four contributing flux terms. The air parcel absorbs part of the LWIR flux from above and below. It is also emitting LWIR radiation upwards and downwards. This emission depends on the local temperature and IR species concentrations. As the air parcel changes altitude, particularly during convective ascent, the temperature change from expansion/compression is generally much larger than the LWIR cooling rate. As an air parcel rises and cools, internal molecular energy is converted to gravitational potential energy. For an ascent rate of 1 km per hour at a lapse rate of -6.5 K km-1, the cooling rate is 6.5 K per hour. From Figure 12, the tropospheric cooling rate from LWIR emission is near 2 K per day or 0.08 K per hour. The rate of cooling during convective ascent may easily be 100 times larger than that produced by LWIR emission. The energy transfer processes related to the tropospheric heat engine and for an air parcel in the troposphere are illustrated schematically in Figure 15.
Figure 15: a) The tropospheric heat engine and b) energy transfer processes for a local tropospheric air parcel (in a plane-parallel atmosphere).
Based on this discussion, the decrease in LWIR flux at TOA produced by an increase in the atmospheric CO2 concentration is the result of absorption within the CO2 bands at lower levels in the atmosphere. The maximum increase in the rate of heating of the troposphere from a ‘doubling’ of the CO2 concentration is less than 0.1 K per day. This is dissipated by the normal convective and radiative energy transfer processes in the troposphere and produces a small increase in both the broadband LWIR emission and the gravitational potential energy as illustrated in Figure 16. Any increase in downward LWIR emission is decoupled from the surface by molecular line broadening effects [Clark, 2013]. In addition, there is a net cooling produced in the stratosphere. This has to be considered as part of the normal diurnal and seasonal stratospheric energy transfer processes. None of the CO2 induced LWIR flux changes in the stratosphere can couple to the surface because of line broadening effects in the lower troposphere. For reference, at a lapse rate of -6.5 C km-1, an increase in temperature of +0.08 C corresponds to a decrease in altitude of 12 meters. This is equivalent to riding an elevator down about four floors.
Figure 16: The dissipation of the ‘radiative forcing’ from a ‘CO2 doubling’ by the normal tropospheric energy transfer processes (schematic). The wavelength specific increase in absorption in the CO2 P and R bands is dissipated as small changes in broadband LWIR emission and gravitational potential energy.
3.2 The Increase in LWIR Flux at the Surface
The IPCC climate assessment reports consider the decrease in LWIR flux at TOA as part of the radiative forcing discussion. In addition to the decrease in upward flux within the CO2 bands, there is also a slight increase in the downward LWIR flux reaching the surface. All of this increase originates from within the first 100 m air layer closest to the surface. There is also a decrease in the downward LWIR flux to the surface from the 200 to 500 m layers immediately above. Figure 17a shows the contribution to the downward flux at the surface from each of the first ten 100 m layers up to 1 km altitude at CO2 concentrations of 280, 400, 560 and 760 ppm. Figure 17b shows the corresponding changes in flux relative to 280 ppm.
Figure 17: Downward LWIR flux to the surface for the first ten 100 m air layers up to 1 km in altitude, a) for CO2 concentrations of 280, 400, 560 and 760 ppm and b) changes in flux relative to 280 ppm. HITRAN radiative transfer calculations, Clark, 2020, 288 K surface and air temperatures, 80% RH.
As discussed above, any small increase in downward LWIR flux to the ocean surface is dissipated as part of the combined LWIR cooling flux and wind driven evaporation (latent heat flux). Over land, the surface temperature is reset each day by the convection transition temperature. The surface and surface air temperatures continue to cool through the night by net LWIR emission from the surface. The temperatures then rise and fall during the day in response to the solar insolation. The effect of an increase in CO2 concentration on the diurnal temperature cycle was investigated by using the minimum air temperatures from the Grasslands 2008 data set [Clark, 2013] as convection transition temperatures in a simple thermal model of the diurnal cycle. The various flux terms were coupled into a surface layer 1 cm thick. The surface layer was coupled to a subsurface thermal conduction model with 200 x 1 cm layers constructed following the finite element method described by Billo [2007]. The thermal properties of dry sand were used for all layers. The LWIR window transmission, the convection coefficient and the rate of air heating were adjusted until the model output was similar to the Grasslands temperature data. The initial LWIR window flux was 46 W m-2. The model was then re-run with the LWIR window flux decreased by 2, 3.7, 5, 10 and 20 W m-2. This represents an increase in CO2 concentration from 280 to approximately 400, 560, 760, 1500 and 8500 ppm [Hansen et al, 2005]. The model was run for full sun ‘clear sky’ conditions and latent heat effects were not included. The daily maximum and minimum surface and surface air temperatures were extracted from the model and the annual averages were calculated. The measured and calculated temperatures using 46 W m-2 for the LWIR flux term are shown in Figure 18 and the average annual temperature changes are shown in Figure 19.
The calculated annual average temperatures are shown in Figure 19a. The changes in temperature as the downward LWIR flux is increased from 280 ppm are shown in Figure 19b. In the model, these are decreases in LWIR window flux from 46 W m-2. The changes in temperature vs. CO2 concentration are shown in Figure 19c and the same data plotted with a logarithmic concentration scale are shown in Figure 19d. The maximum observed increase is in the minimum surface temperature. However, for a conventional CO2 ‘doubling’ to 560 ppm, this increase is only 0.3 C. Even at a CO2 concentration of approximately 8500 ppm, the increase in surface temperature is only 1.3 C. The increases in surface air temperature are less. The increase in air temperature for a CO2 doubling is less than 0.1 C. At 8500 ppm, the air temperature increase is near 0.4 C. This demonstrates that the radiative forcing/climate sensitivity temperature estimates used in the Paris accord have no basis in physical reality.
Figure 18: Calculated and measured Grassland min/max surface and air temperatures.
Figure 19: Calculated changes in 2008 Grassland annual average temperatures a) temperature vs increase in downward LWIR flux, b) increases in temperature from a) with 280 ppm baseline subtracted, c) data from b) plotted vs. increase in CO2 concentration and d) data from c) plotted with concentration on a logarithmic scale.
3.3 Water Vapor Feedback
The concept of water vapor feedback is a mathematical artifact that was created by the original modeling assumptions used by Manabe and Wetherald [M&W, 1967]. They assumed an exact flux balance between a 24 hour average solar flux and the average OLR. The surface was a partially reflective blackbody surface with zero heat capacity and the relative humidity distribution was fixed. By definition within this oversimplified model, an increase in atmospheric CO2 concentration must create an increase in surface temperature to restore the flux balance at TOA. Furthermore, as the temperature increases, there must be an increase in water vapor concentration to maintain the fixed RH assumption. This provides the ‘feedback’.
In reality, the humidity depends on the local weather system. If the absolute water vapor concentration is approximately constant within a given air mass, the RH will decrease during the day as the air is heated by convection from the solar heated surface. If there is local heating from air compression produced within ‘blocking’ high pressure systems or downslope winds, then increases in temperature are associated with a decrease in relative and absolute humidity. Changes in humidity are illustrated in Figure 20. This shows the air temperature, relative and absolute humidity for days 120 to 180 from the Grasslands data set associated with Figure 18 [Clark, 2013]. The areas highlighted in yellow are offshore flow conditions where the air is heated by downslope flow. The humidity decreases and the air temperature increases. This is the opposite of the ‘feedback’ assumed by M&W. The areas highlighted in green are onshore flow conditions with relatively constant absolute humidity in the 12 to 14 mmol mol-1 range. The recorded diurnal change in RH is from approximately 60 to 94%. (The upper limit to the RH measurement is 94%).
Approximately half of the downward LWIR flux emitted to the surface originates from within the first 100 m layer above the surface. Changes in LWIR flux related to humidity effects are dominated by changes in air temperature and water vapor concentration close to the surface. This is also part of the turbulent surface boundary layer [Garai and Kleissl, 2011].
Figure 20: Changes in a) air temperature, b) relative humidity (%) and c) absolute humidity (mmol mol-1) for days 120 to 180 from the 2008 ‘Grasslands’ data set. There is no simple ‘feedback’ relationship between RH and temperature.
There is also another important underlying issue in the M&W feedback assumption. This is the separation of the LWIR flux from the latent heat flux. These two are fully coupled. The water molecules that evaporate from the surface are transported to higher altitudes as part of the convective mass transport. They are coupled into a thermal and humidity gradient. Over the oceans, the latent heat flux depends on the humidity gradient and the wind speed. The combined net LWIR and latent heat flux Qirnet + Qlat = Qcool can be calculated using:
Here, sigma is Stefan’s constant, Qirwin is the IR transmission window flux, klat is an empirical coupling coefficient, PTws is the saturated water vapor concentration at the surface temperature Ts, PTwa is the saturated water vapor concentration at the surface air temperature Ta, Rh is the relative humidity and U is the wind speed [Clark, 2013, Yu et al, 2008]. There is also a small sensible heat flux term that is not included here.
When the atmospheric CO2 concentration is increased, the increase in downward LWIR flux is coupled initially into the first 100 micron layer of the ocean surface. There is a small increase in surface temperature that increases both the net LWIR flux and the latent heat flux. In practice, these changes are obscured by the much larger variations in wind driven latent heat flux. However, to illustrate the error in the feedback assumption, consider the hypothetical case of constant wind speed. If there are no other changes in the heat transfer, the total increase in cooling flux Qcool should match the increase in the downward LWIR flux. This can be estimated from Eqn. 2. As the ocean surface temperature (SST) increases, the latent heat flux Qlat increases faster than the increase in blackbody radiation and the required change in temperature decreases. Representative changes in temperature and the related changes in blackbody radiation and latent heat flux are shown in Figure 21 for increases in downward CO2 flux to the ocean of 2, 4 and 8 W m-2. These correspond approximately to the observed increase of 120 ppm, a doubling and a quadrupling of the atmospheric CO2 concentration. Here, the air temperature was set 1 C below the surface temperature, the RH was 75%, the wind speed was 6 m s-1 and the coupling coefficient was 2.2. From Figure 21a, the increases in ocean temperature at 15 C SST are 0.23, 0.46 and 0.92 C for flux increases of 2, 4 and 8 W m-2. This is an ideal case in which only the temperature and the temperature dependent vapor pressure terms are changed in Eqn. 2. However, this illustrates the combined role of the net LWIR flux and the latent heat flux in changing the surface temperature. For reference, the increase in temperature for a blackbody surface without any evaporation are shown as the dotted lines in Figure 21a. At 15 C, the latent heat flux reduces the blackbody heating effect by approximately 40%. This is the opposite of the feedback effects used in the IPCC climate models to ‘amplify’ the presumed warming from the increase in the LWIR flux from an increase in the atmospheric CO2 concentration.
Figure 21: a) Increase in ocean surface temperature required to counteract an increase in downward LWIR flux from CO2 of 2, 4 and 8 W m-2 for the ideal case of constant wind speed. The solid lines show the change in temperature estimated from Eqn. 2. The dashed lines show the temperature increase for a blackbody surface without any evaporation. b) The changes in blackbody emission and latent heat flux produced by the temperature increases from Eqn. 2 shown in a). The air temperature was set 1 C below the surface temperature, the RH was 75%, the wind speed was 6 m s-1 and the coupling coefficient was 2.2.
4.0 CLIMATE SENSITIVITY
In order to determine the change in surface temperature from the change in ‘radiative forcing’, the basic logic of cause and effect has been ignored and replaced by correlation. It has been assumed a-priori, that all of the observed increase in the surface temperature record has been caused by an increase in radiative forcing, produced mainly by the increase in atmospheric CO2 concentration. This has been used to create a ‘climate sensitivity’ to CO2. In reality, none of the observed surface temperature increase can be caused by CO2. The increase downward LWIR flux to the surface produced by the measured 120 ppm increase in atmospheric CO2 concentration cannot cause any measurable temperature change. Instead, the temperature increases can be explained by a combination of the influence of the Atlantic Multi-decadal Oscillation (AMO) on the land surface air temperatures, urban heat island effects, changes in the number and location of the weather stations and the ‘adjustments’ used to create the global average temperature record from the raw weather station data. In order to understand the pseudoscientific creation of the climate sensitivity, it is necessary to consider the ocean gyre circulation and the related ocean oscillations, the coupling of these ocean oscillations to the weather station record and the creation of the global average temperature anomaly from the weather station data. These areas will now be considered in more detail.
4.1 The ocean Gyre Circulation and the Ocean Oscillations
The downward LWIR flux from the lower troposphere creates the so called ‘greenhouse effect’ by ‘blocking’ most of the upward LWIR flux from the surface. This produces the LWIR exchange energy that limits the net LWIR cooling flux. The excess surface heat is removed by moist convection or evapotranspiration, which is a mass transport process. The coupling of the convection to the rotation of the earth leads to the formation of the trade winds that drive the ocean gyre circulation. There is no requirement for an exact local flux balance at the ocean surface between the solar heating and the surface cooling. This leads to natural oscillations in ocean surface temperature. There is no simple mathematical solution to the fluid dynamics of the air-ocean interface and the oscillations are quasi-periodic. The oceans and their fluid dynamics representations can shift from one state to another. The ocean gyre circulation and the four main ocean oscillations are shown in Figure 22 [AMO, 2020, ENSO, 2020, IOD, 2020, PDO, 2020].
Figure 22: The ocean gyre circulation and the four main ocean oscillations.
There are five main ocean gyres located in the N. and S. Atlantic, the N. and S. Pacific Oceans and the S. Indian Ocean. Each gyre has its own unique circulation properties. The three southern ocean gyres are coupled to the Southern Ocean circulation. The equatorial circulation in the Atlantic and Pacific Oceans is shifted several degrees north of the equator. The S. Atlantic Equatorial Current splits off the coast of Brazil and part feeds into the N. Equatorial Atlantic Current and the Caribbean Current. The ENSO and the IOD have periods between 3 and 10 years. The PDO varies on both a 10 to 25 and a 60 to 70 year time scale. The AMO has a period in the 60 to 70 year range. The peak ocean surface temperature is near 30 C in the equatorial warm pools. The equatorial ocean oscillations involve changes in the area and location of these warm pools.
4.2 The Global Average Surface Temperature Anomaly and the AMO
In order to evaluate the output of the climate models, the recorded weather station data has been processed into a ‘global average surface temperature anomaly’. This is an area weighted average of the weather station data after it has been extensively processed or ‘homogenized’ and the mean has been subtracted. These adjustments are supposed to remove bias and account for missing data. They have also added significant warming to the raw data. This has been considered in more detail by Andrews [2017a, 2017b and 2017c] and by D’Aleo [2010]. When the climate anomaly record, such as the HasdCRUT4 data set is evaluated, the dominant term is found to be the Atlantic Multi-decadal Oscillation (AMO). The correlation coefficient between the two data sets is 0.8. This is illustrated in Figure 23a. The influence of the AMO extends over large areas of N. America, Western Europe and parts of Africa. The weather systems that form over the oceans and move overland couple the ocean surface temperature to the weather station data through the diurnal convection transition temperature. The contributions of the other ocean oscillations to the global temperature anomaly are smaller. The IOD and the PDO are dipoles that tend to cancel and the ENSO is limited to a relatively small area of the tropical Pacific Ocean. However, small surface temperature variations in the tropical oceans have a major impact on ocean evaporation and rainfall. Figure 23b shows a tree ring construction of the AMO from 1567 [Gray, 2004, Gray.NOAA, 2021]. The modern instrument record is also indicated in green. None of the temperature changes related to the AMO can be attributed to an increase in atmospheric CO2 concentration.
Figure 23: a) Plots of the HadCRUT4 and AMO temperature anomalies overlapped to show the similarities. Both the long term 60 year oscillation and the shorter term ‘fingerprint’ details can be seen in both plots. The role of ‘adjustments’ in the 0.3 C offset since 1970 requires further investigation. b) Tree ring reconstruction of the AMO from 1567.
4.3 The Creation of the ‘Climate Sensitivity’
The HadCRUT4 global average temperature anomaly record and other similar data sets have been used to create two pseudoscientific ‘climate sensitivities’. The first is an ‘equilibrium climate sensitivity’ (ECS) and the second is a ‘transient climate response’ (TCR). The ECS is the equilibrium climate temperature response to a ‘CO2 doubling’ after the model oceans have adjusted to a new equilibrium state and the TCR is the response to a gradual increase in the radiative forcing, usually from a 1% per year increase in CO2 concentration before equilibrium is reached [IPCC 2013 Chapter 9].
Otto et al [2013] define these as:
Here, F2x is the radiative forcing produced a doubling of the atmospheric CO2 concentration, set in this case to 3.44 W m-2 for a doubling from ‘preindustrial levels’, 280 to 560 ppm [Meinhausen et al, 2011], The change in temperature is taken from the HadCRUT4 global temperature anomaly [HadCRUT4, 2019] and the radiative forcings are taken from the CMIP5 /RCP4.5 model ensemble. The change in heat content is dominated by ocean heat uptake. More recent estimates of ECS and TCR are provided by Lewis and Curry [2018]. The decadal temperature and forcing estimates from data given by Otto et al [2013] are shown in Figures 24a and 24b. The 1910 AMO cycle minimum and the 1940 maximum are indicated. The increase in the downward LWIR flux related to the ‘radiative forcing’ shown in Figure 24b cannot couple below the ocean surface and cause any measurable change in ocean temperature.
Figure 24: a) Decadal mean temperature estimates derived from the HadCRUT4 global mean temperature series b) decadal mean forcing with standard errors from the CMIP5 /RCP4.5 ensemble. Data from Otto et al [2013].
Using the data from Figure 24 and estimates of Q from various sources, Otto et al assume that their net radiative forcing estimates are responsible for the observed heating effects and that the temperature response to the change in LWIR flux is linear. Plots of DT vs (DF-DQ) and DT vs DF are therefore presumed to be linear with a slope that changes with the value of ECS or TCR. The results generated by Otto et al are shown in Figure 25. Using the data for 2000 to 2010, they give an ECS of 2.0 C with a 5-95% confidence interval of 1.2 to 3.9 C and a TCS of 1.3 C with a confidence level of 0.9 to 2.0 C.
Figure 25: Estimates of ECS and TCR from Otto et al [2013]
It is claimed that the ECS for the CMIP5 model ‘ensemble’ is in the range from 2.1 to 4.7 C. In the US, this modeling effort is coordinated by the climate group at Lawrence Livermore National Laboratories (LLNL). They also maintain the ‘library’ of climate model results [Stauffer et al, 2017, Taylor et al, 2012]. The CMIP5 model results were used by the UN Intergovernmental Panel on Climate Change (IPCC) in their fifth Climate Assessment Report (AR5) [IPCC 2013, Chap. 9]. For the AR6 IPCC report, the ECS range of the CMIP6 climate model ‘ensemble’ is given as 1.8 to 5.6 K [Hausfather, 2019]. These climate sensitivities are shown in Figure 26. The median ECS of 3.8 C/280 ppm translates into a temperature sensitivity of about 74 ppm C-1. A 2 C temperature rise corresponds to a CO2 concentration of approximately 430 ppm. This is the pseudoscientific basis of the 2 C temperature limit incorporated into the Paris Climate Accord [Luning and Vahrenholt, 2017]. The radiative forcings published in the latest IPCC climate assessment report [IPCC AR6 Wgp 1 Chap 7, 2021] are shown in Figure 27 and the changes in temperature over time derived from these forcings are shown in Figure 28. The forcings for the ‘greenhouse gases’ are derived from radiative transfer calculations using the HITRAN database or similar data. The various aerosol terms are simply ‘tuning knobs’ that can be adjusted to give a better fit to the measured temperature data.
Figure 26: Pseudoscientific equilibrium climate sensitivity (ECS) for a doubling of the CO2 concentration from 280 to 560 ppm for selected CMIP5 and CMIP6 climate models.
Figure 27: ‘Effective’ radiative forcings from AR6 WGp 1 Chapter 7.
Figure 28: Temperature changes from 1750 attributed to the ‘forcings’ shown in
Figure 27.
The role of the AMO in setting the surface air temperature has been misunderstood or ignored for a long time. The first person to claim a measurable warming from an increase in CO2 concentration was Callendar in 1938. He used weather station temperatures up to 1935 that included most of the 1910 to 1940 warming phase of the AMO [Callendar, 1938]. The warming that he observed was from the AMO not CO2. During the 1970s there was a ‘global cooling’ scare that was based on the cooling phase of the AMO from 1940 to 1970 [McFarlane, 2018, Peterson et al, 2008, Douglas, 1975, Bryson and Dittberner, 1976]. In their 1981 paper Hansen et al chose to ignore the 1940 AMO peak in their analysis of the effects of CO2 on the weather station record [Hansen, 1981]. Similarly Jones et al conveniently overlooked the 1940 AMO peak when they started to ramp up the modern global warming scare in 1986 [Jones et al, 1986]. This is illustrated in Figure 29. The AMO and the periods of record used are shown in Figure 29a. The temperature records used by Callendar, Douglas, Jones et al and Hansen et al are shown in Figures 29b through 29e. The Keeling curve showing the increase in atmospheric CO2 concentration is also shown in Figures 29d and 29e [Keeling, 2021].
Figure 29: a) AMO anomaly and HadCRUT4 global temperature anomaly, aligned from 1860 to 1970, b) temperature anomaly for N. temperate stations from Callendar [1938], c) global cooling from Douglas [1975], d) global temperature anomaly from Hansen et al, [1981] and e) global temperature anomaly from Jones et al, [1986]. The changes in CO2 concentration (Keeling curve) are also shown in c and d. The periods of record for the weather station data are also indicated.
5.0 LORENZ INSTABILITIES
In 1963, Lorenz was evaluating a simple model of three coupled non-linear equations that described two dimensional convective flow. He discovered that the solutions to these equations were unstable and sensitive to small changes in the starting variables [Lorenz, 1963]. This is illustrated in Figure 30. He later established that weather forecasting predictions were limited to approximately 12 days ahead before the model instabilities became dominant [Lorenz, 1973].
Figure 30: Lorenz instabilities for 3 coupled partial differential equations, from Lorenz, 1963.
The large scale climate models used today require the solution of a very large number of coupled non-linear equations. There is no reason to expect such models to have any predictive capabilities. They are inherently unstable and have to be ‘tuned’ to give the desired results. In fact, they can be ‘tuned’ to give any desired result. All of the climate models shown in Figure 26 have been tuned based on the pseudoscience of radiative forcing to create the global surface temperature anomaly. The fact that all of these models have a similar climate sensitivity to CO2 is clear evidence of the climate fraud. A realistic climate model should not show any ‘climate sensitivity’ to CO2. The basic requirement of any climate model is that it should predict the measured variables of the climate system. This means the measured minimum weather station temperature and the delta T or change from minimum to maximum MSAT. Any weather station bias terms should be incorporated into the model. The measured data should not be changed. Since the dominant term in the climate record is the AMO, the climate models should be capable of predicting the ocean oscillations and global ocean temperatures derived from these oscillations. Figure 31 shows global ocean temperatures from 1979 to 2021 and 68 model ‘predictions’ from 13 different CMIP6 models [Spencer, 2021]. The models have clearly failed. The ‘predicted’ ocean surface temperatures are too large. This is to be expected, based on consideration of Lorenz instabilities and model ‘tuning’. Such models may perhaps best be described as quasi-stable pseudo-random number generators, all tuned to the same temperature series. The pseudo-random number generators used in Monte Carlo calculations give exactly the same random number sequence for each model run until the seeds used in the generator are changed. Here, the models are unstable because of the underlying Lorenz instabilities related to rounding errors and other modeling effects. Two model runs for exactly the same model conditions run on the same computer will give different results. The spread between the maximum and minimum model results increases from approximately 0.6 C to 1.6 C over time span of the model runs. This is characteristic of Lorenz instabilities.
Figure 31: Global sea surface temperatures (60° N to 60°S), 1979 to 2021, CMIP6 models compared to ERSSTv5 observations (thick black line) [Spencer, 2021]
6.0 THE EVOLUTION OF THE CLIMATE FRAUD
No one person created the created the climate fraud. It evolved over a long period of time. There are at least three different parts to this. First, melodramatic prophecies of the global warming apocalypse became such a good source of research funding that the scientific process of hypothesis and discovery collapsed. Research also became more expensive. Isotope ratio studies required advanced mass spectrometers. The climate models needed the largest and fastest computers available, and the personnel to operate them. Second, there was institutional fraud related to ‘mission creep’ within various government agencies. NASA was established to put a man on the moon. There was no provision to shut it down after that mission was accomplished. Similarly, the Atomic Energy Commission was established to develop nuclear energy for military and commercial applications. As resources diminished, the government scientists looked for other jobs. Some created them using the climate apocalypse. Third, there was a deliberated decision by various outside interests, including environmentalists and politicians to exploit the climate apocalypse to further their own causes. Some environmentalists wanted to disrupt the energy supply to reduce human population. Various left wing political groups wanted to disrupt the energy supply to promote their revolutionary beliefs. More mainstream politicians then realized they could use the climate apocalypse to increase taxes and buy votes. Gradually a vast secondary industry of policy analysts, economists, geologists, geographers, ecologists, psychologists, sociologists and other assorted ‘experts’ was created and funded to study every aspect of this nonexistent global warming apocalypse problem. All of this is a massive pyramid or Ponzi scheme built on the invalid hypothesis of an equilibrium average climate. The peer review process in climate science has collapsed and been replaced by blatant cronyism. The climate modelers have retreated into a cocoon of lies where they discuss the pseudoscience of radiative forcings, feedbacks and climate sensitivities to a CO2 ‘doubling’. This is just climate theology. How does a doubling of the atmospheric CO2 concentration change the number of angels that may dance on the head of a climate pin? They are playing computer games in an equilibrium average climate fantasy land. They use their ‘advanced’ climate models to create the sacred spaghetti plots of global warming. This is GIGO: garbage in, gospel out. The model ‘predictions’ are fed directly to government agencies and the IPCC with minimal outside scrutiny. There is no climate science involved. Irrational belief in computer models has replaced the Laws of Physics. The Imperial Cult of the Global Warming Apocalypse has claimed the Divine Right to save the world from a non-existent problem.
6.1 The Transition from Invalid Hypothesis to Climate Fraud
There is no single person or single event that can be identified as the source of the climate fraud. There is no ‘smoking gun’. One of issues that needs to be established is the date or dates of the transition from invalid hypothesis to downright fraud. This is a similar concept to that of ‘obviousness’ in Patent Law. How would a scientist or engineer verify that a climate model was giving a valid result? When does the fraud become obvious?
A timeline of selected developments in climate science is presented in Figure 32 and the events indicated are listed in Table 1. The AMO, including the oscillation and the linear slope (blue), and the increase in atmospheric CO2 concentration (Keeling curve, green) are also shown [AMO, 2020, Keeling, 2020]. The transition from hypothesis to fraud is indicated. It took place over the 12 year period from the publication of the radiative convective equilibrium model by Manabe and Wetherald in 1967 to the publication of the Charney Report in 1979. The fraud certainly becomes obvious with the 1981 Science publication by Hansen et al. These events will now be considered in more detail.
Figure 32: Timeline of climate developments
Table 1: List of climate developments shown in Figure 32.
Fourier, in his ‘memoires’ of 1824 and 1827 was interested in the temperature of the earth. His theory of heat, published in 1822 was successful in explaining detailed observations including seasonal changes and the time delays or phase shifts in the subsurface temperature [Fourier, 1822, 1824, 1827].
At a moderate depth, as three or four meters, the temperature observed does not vary during each day, but the change is very perceptible in the course of a year, it varies and falls alternately. The extent of these variations, that is, the difference between the maximum and minimum of temperature, is not the same at all depths, it is inversely as the distance from the surface. The different points of the same vertical line do not arrive at the same time at the extreme temperatures. The extent of the variations, the times of the year, which correspond to the greatest, to the mean, or to the least temperatures, change with the position of the point in the vertical line. There are the same quantities of heat which fall and rise alternately, all these values have a fixed relation between themselves, which are indicated by experiments and expressed distinctly by the analysis. The results observed are in accordance with those furnished by the theory, no phenomenon is more completely explained.
Fourier 1824
The equilibrium hypothesis was first proposed by Pouillet in 1836. He was apparently unaware of the implications of Fourier’s work on the subsurface temperature phase shift [Pouillet, 1836]. The time delay or phase shift between the peak solar flux and the temperature response provides irrefutable evidence of a non-equilibrium climate. Such seasonal and diurnal phase shifts have been ignored in climate science for almost 200 years. The most easily observed phase shift is the 4 to 8 week delay between the peak solar flux at solstice and the seasonal temperature response at mid and higher latitudes, illustrated above in Figures 6 and 7. This has been recorded part of normal weather station temperature data for well over 100 years.
Once the work of Agassiz [1840] on the existence of an Ice Age was accepted, after more than 20 years of rather acrimonious debate, the climate discussion changed [Imbrie and Imbrie, 1979]. It was no longer about the heating and cooling of the earth. Instead it was about a mechanism that could cause an Ice Age. This led Tyndall [1861, 1863] to speculate that changes in the atmospheric CO2 concentration could alter the earth’s climate and this in turn was the motivation for Arrhenius [1896] to try and calculate a possible change in temperature produced by CO2. He was interested in both long term heating and cooling. Arrhenius had no temperature measurements that he could use to validate his calculations. His results were invalid because he used the equilibrium climate assumption. However, the idea that an increase in CO2 concentration could lead to an increase in surface temperature had been established. The first person to claim a measurable effect on surface temperature from an increase in CO2 concentration due to combustion was Callendar [1938]. He assumed that an increase in LWIR absorption and emission in the CO2 band near 670 cm-1 could cause a change in surface temperature. He found a slight increase in both CO2 concentration and meteorological temperatures, particularly in the N. hemisphere. He was probably the first person to find the AMO signal in the weather station data, since his period of record included the warming phase of the AMO from about 1910 to 1935. This is illustrated above in Figure 29.
In the mid-1950s, improved spectroscopic measurements and line by line computer calculations allowed Plass [1956a] to provide improved estimates of possible heating effects from CO2. He calculated a cooling rate for CO2 of 0.2 to 0.3 K per day in the troposphere. He was still using the equilibrium assumption and estimated changes in surface temperature of +3.6 C and -3.8 C for a doubling and a halving of the CO2 concentration from 330 ppm. In a different paper, he discussed a ‘CO2 Theory of Climate Change’ [Plass, 1956b]. Here it is clear that he was interested mainly in changes in CO2 concentration related to carbonate rock weathering as a cause of an Ice Age cycle, although fossil fuel combustion was also discussed. He speculated the when all of the known coal and oil reserves were used up in less than 1000 years, the equilibrium climate temperature rise could be 12 C with the CO2 concentration increasing to 3000 ppm.
Interest in the effects of CO2 from fossil fuel combustion on climate was revived in the late 1950s with the work of Burt Bolin and Roger Revelle on the distribution of CO2 between the atmosphere and the oceans [Bolin and Eriksson, 1959, Bolin, 1960, Revelle and Seuss, 1957]. They had a new technique that they could use. This was the measurement of the carbon isotope ratios 14C/12C and 13C/12C. However, this only provided information on the amount of CO2 in the atmosphere that could be attributed to combustion. There was no new information on the relationship between CO2 and surface temperature. They also used exaggerated claims of climate warming to obtain research funds. They made no attempt to validate their claims using any thermal engineering calculations of the surface temperature.
In order to develop a computer based climate model, it was necessary to develop suitable radiative transfer algorithms that could simulate the changes in energy transfer in the atmosphere as the concentration of such IR active species as H2O and CO2 changed. All of this work relied on the equilibrium assumption. Early efforts were described by Manabe and Moller [1961]. In a review of atmospheric optics Moller [1964] used the equilibrium assumption to derive an invalid ‘effective radiation temperature’ of 250 K, using 0.35 for the planetary albedo or reflectivity of the earth. He also described the troposphere as a heat engine, although he did not elaborate on this. The first generally accepted radiative transfer climate model, published by Manabe and Wetherald (M&W) [1967], was also based on the equilibrium climate assumption. It was really just a mathematical platform for the development and evaluation of radiative transfer and related algorithms. The assumptions had to create climate warming even before the model code was written.
6.3 Invalid Hypothesis to Climate Fraud: From Manabe and Wetherald to the Charney Report, 1967 to 1979
The assumptions used by M&W in the development of their 1967 model were clearly and honestly stated on the second page of their paper:
1) At the top of the atmosphere, the net incoming solar radiation should be equal to the net outgoing long wave radiation.
2) No temperature discontinuity should exist
3) Free and forced convection and mixing by the large scale eddies prevent the lapse rate from exceeding a critical lapse rate equal to 6.5 C km-1.
4) Whenever the lapse rate is subcritical, the condition of local radiative equilibrium is satisfied.
5) The heat capacity of the earth’s surface is zero.
6) The atmosphere maintains the given vertical distribution of relative humidity (new requirement).
These assumptions contain three fundamental scientific errors. 1) There is no exact flux balance at TOA, 2) the heat capacity and the effects of moist convection have to be included in the surface heating and 3) the relative humidity distribution is not fixed. In addition, molecular line broadening in the lower troposphere means that the upward and downward LWIR flux are not equivalent. The use of such a simple model as a mathematical platform for the development of radiative transfer algorithms is entirely reasonable, provided that the limitations are understood and clearly stated. Any scientist who uses a climate model that incorporates the M&W assumptions should add the caveat ‘does not apply to planet earth’. Has anyone seen a ‘24 hour average sun’ shining in the sky at night? Unfortunately, the climate warming artifacts created by the M&W model soon became a lucrative source of research funds and the limitations of the M&W model were conveniently overlooked. This is still the case today. The original M&W model was recently re-evaluated using an updated code [Kluft, 2020]. All of the original M&W assumptions were accepted without question.
The M&W model created two ‘bandwagons’ that could be used to obtain research funding. First, the radiative transfer algorithms could be improved with better spectroscopic constants and more greenhouse gases. Second, the M&W model could be incorporated into a general circulation model (GCM) with well over a thousand M&W ‘units’ coupled together within a modified weather forecasting program to make ‘improved’ climate ‘predictions’. None of this required any change to the underlying M&W ‘model’ assumptions. Within a decade, an additional 11 ‘minor’ species had been added to the M&W model including CH4, NH3, N2O, HNO3, SO2 and six halogenated hydrocarbons [Ramanathan and Coakley (R&C), 1978]. All of the improvements in the IR spectroscopy were valid, until they were used to calculate the ‘equilibrium temperature’ in the M&W model. For example, melodramatic claims of “An appreciable increase in global surface temperature” were made for halogenated hydrocarbons without out any quantitative thermal analysis [Ramanathan, 1975]. All of this was still based on ‘radiative convective equilibrium models’. This was unambiguously stated by R&C on page 479 of their 1978 review paper.
“For radiative-convective equilibrium the net outgoing longwave radiative flux at the top of the atmosphere Fn0 must equal the net solar radiative flux Sn0. Likewise, because the stratosphere is in radiative equilibrium, Fn1 must equal the net solar radiative flux into the troposphere Sn1. For any perturbation, the stratosphere and atmosphere as a whole seek a new state of radiative equilibrium.”
In order to develop a so called general circulation model (GCM) for climate simulation, the M&W model had to be incorporated into the fluid dynamics equations used to describe atmospheric and ocean flow. These are a complex set of complex partial differential equations that have to be solved numerically. There is no analytical solution. Lorenz [1963] showed that the solutions for such equations were inherently unstable. In weather forecasting, this means that the GCM solutions can be accurate for projections up to about 10 days ahead. The accuracy of a weather forecast is easy to establish by comparison of prediction to measurement. There is no reason to expect a complex GCM climate model to have any predictive capability over the time scales required for climate change because of these instabilities. This is discussed above in Section 5. The equilibrium assumption was used to impose an invalid ‘solution’ to the large numbers of coupled non-linear equations in the GCMs. The climate models were run until a stable solution was obtained where the planetary average LWIR flux matched the planetary average absorbed solar flux. By 1975, a ‘highly simplified’ GCM had been developed by M&W [1975]. They chose to ignore the assumptions they made in 1967 and continued with the development of an invalid climate model. The 1967 ‘model’ was now described as a ‘global average climate model’. Although the M&W GCM did not contain any real climate effects such as ocean transport and the cloud cover was fixed, claims of global warming from a ‘CO2 doubling’ were still made.
The M&W approach was officially ‘sanctified’ by the Charney report [1979]. This was a review of CO2 induced warming effects derived from equilibrium climate models, including feedback effects. At the time of the review, the results from only five GCMs were available, 3 from Manabe’s group and 2 from Hansen’s group. The reviewers concluded that a warming of 3±1.5 C from a ‘doubling’ of the atmospheric CO2 concentration was likely. The mathematics used in the climate ‘models’ appeared reasonable based on the acceptance of the invalid equilibrium assumption, so no further investigation was needed. Lorenz’s work and the limitations of the weather forecasting models were ignored.
The reviewers involved in the Charney report also chose to ignore the history of CO2 induced climate change and its origin as the explanation of the cause of an Ice Age cycle. The real cause of an Ice Age, changes to the eccentricity of the earth’s orbit, had been established from an analysis of ocean sediment cores in 1976 [Hays et al, 1976]. A more detailed description was given in the book ‘Ice Ages’ by Imbrie and Imbrie [1979]. Since changes in CO2 concentration did not cause an Ice Age, there was no reason to expect that such changes from fossil fuel combustion would cause climate change. Tyndall’s speculations from the 1860’s had been disproved. However, the bandwagons were rolling and the global warming apocalypse had already been declared based on the mathematical artifacts created by the equilibrium climate models.
6.4 Hansen’s 1981 Additions to the Climate Modeling Fraud
By 1979 sufficient evidence had accumulated to show that there was no reason to expect that changes in atmospheric CO2 concentration could cause any climate change. Detailed data showing the non-equilibrium response of land surface temperatures were available a decade before M&W with the publication of ‘Exploring the Atmosphere’s First Mile’ by Lettau and Davidson [1957]. Hubert Lamb’s work ‘Climate, Past Present and Future’ was published in 1972. Here he discussed the role of wind driven evaporation in causing long term weather anomalies [Lamb, 1972]. These involved much larger changes in ocean latent heat flux than any change in LWIR flux from CO2. Revised spectroscopic constants for water from 200 nm to 200 µm were published by Hale and Querry in 1973. These showed that the penetration depth of the LWIR radiation from CO2 into water was less than 100 µm [Hale and Querry, 1973]. By the early 1970’s, high quality satellite IR radiometer data were available from the Nimbus program that showed the variability in the LWIR emission to space [Hanel et al, 1971]. As discussed above, CO2 did not cause Ice Ages.
Instead of correcting the underlying errors related to the equilibrium assumption, Hansen et al added three more invalid assumptions [Hansen et al, 1981]. First, a ‘slab’ ocean was added to the M&W model. However, there was no consideration of surface energy transfer effects. In particular, wind driven evaporation was ignored and the LWIR flux was assumed to heat the ocean even though the penetration depth of the LWIR flux was only 100 micron. The ‘ocean’ basically added heat capacity and a time delay to the model. This provided the pseudoscientific foundation for ‘climate sensitivity’. Second, a prescribed mathematical ritual of ‘radiative forcing’ was introduced. This was based on the perturbation of an equilibrium climate state by a ‘doubling’ of the CO2 concentration followed by the transition to a new equilibrium state with a higher surface temperature. Small changes in equilibrium LWIR flux were assumed to be capable of changing the surface temperature including the oceans. No thermal engineering calculations of the change in surface temperature were performed to validate the model. The changes in ‘equilibrium flux’ are shown in Figure 33 and calculated changes in ‘equilibrium temperature’ are shown in Figure 34.
Figure 33: Discussion of the effects of a hypothetical ‘CO2 doubling’ from 300 to 600 ppm on an equilibrium average climate from Hansen et al, 1981.
Figure 34: Effects of various ‘forcing agents’ on surface temperature calculated using an equilibrium average climate model artificially constrained by an exact flux balance at TOA.
Third, there was a ‘bait and switch’ change from ‘equilibrium’ surface and air temperatures to the weather station temperature record. The various flux terms interact with the surface, not the weather station thermometer located in a ventilated enclosure at eye level above the ground [Oke, 2006]. There was no change to the model, the output was just renamed. The weather station record that was presented by Hansen et al also included the well-defined AMO peak near 1940. They chose to ignore this and called it ‘noise’. This is shown above in Figure 29e. Since 1981, the only change to the basic equilibrium climate model was the addition of ‘efficacies’ to the radiative forcings by Hansen et al [2005].
In 2008, Knutti and Hegerl stated:
"When the radiation balance of the Earth is perturbed, the global surface temperature will warm and adjust to a new equilibrium state."
The fraud in Hansen et al should be obvious to a scientist or engineer with a basic understanding of surface energy transfer. The ocean surface energy transfer was ignored, especially the wind driven evaporation. The normal diurnal and seasonal changes in surface flux and temperature are sufficiently large that the small changes in flux from CO2 are too small to have any effect. At present, the average atmospheric concentration of CO2 is increasing by approximately 2.4 ppm per year. This produces an increase in LWIR flux of 0.034 W m-2 per year. The step doubling by 280 ppm is a hypothetical modeling construct.
6.5 The Growth of the Climate Modeling Fraud
When the Charney report was written in 1979, results from global circulation models (GCMs) were only available from two groups. One was associated with Manabe at NOAA and the other with Hansen at NASA Goddard. A total of five rather primitive models had been developed. Much of the data had not been published and was communicated privately. By 1995, 18 coupled climate models were available from seven different countries [Meehl et al, 1997]. The modeling effort for the IPCC is now coordinated through the Coupled Model Intercomparison Project (CMIP). In 2019 there were 49 modeling groups with approximately 100 different models involved in CMIP6 generating the fraudulent data to be incorporated into the next IPCC climate assessment (AR6) [Hausfather, 2019]. All of these models used the same basic approach established by M&W and Hansen. The invalid assumption of an equilibrium average climate is still the foundation of the CO2 induced warming generated by these models. The ‘equilibrium climate sensitivity’ for various CMIP5 and CMIP6 climate models are shown above in Figure 26. As computer technology has improved, the models have become much more complex. A lot of effort has been expended on ‘tuning’ these models to match observational data, but the fundamental instabilities related to the solution of the coupled non-linear equations have been ignored. As new groups have joined the CMIP modeling effort, the underlying assumptions have been accepted without question. There has been no ‘due diligence’ to investigate the validity of the models. The climate modelers have withdrawn into their own fictional modeling world. They are no longer scientists. They have become prophets of the Imperial Cult of the Global Warming Apocalypse. They remain inside a cocoon of lies and discuss the pseudoscientific climate theology of ‘forcings’, ‘feedbacks’ and ‘climate sensitivities’ among themselves. The peer review process has collapsed and been replaced by blatant cronyism. Their primary mission is to perpetuate the climate pyramid scheme by providing the climate lies that governments and special interest groups want to hear and exploit.
6.6 The Political Exploitation of the Climate Modeling Fraud
The political exploitation of the climate modeling fraud started in the 1970s over exaggerated concerns related to population growth. Paul Ehrlich published his book ‘The Population Bomb’ in 1968 and Meadows et al published ‘Limits to Growth’ in 1972 [Meadows et al, 1972]. An important event was the 1975 conference ‘The Atmosphere Endangered and Endangering’ organized by anthropologist Margaret Mead [Hecht, 2007, Mead and Kellogg, 1976]. Her objective was to exploit atmospheric pollution - real or imagined - for population control. Attendees included Stephen Schneider and John Holdren. Both were strongly influenced by Ehrlich. Schneider became a leading advocate of the CO2 climate scare at Stanford University. Holdren became science and technology advisor to President Obama.
Efforts also started to exploit global warming within the World Meteorological Organization (WMO) and the United Nations Environmental Program (UNEP). Bert Bolin conducted climate research for both agencies. Maurice Strong was the first head of UNEP in 1972 and from the start it was involved in blatant environmental advocacy [McClean, 2009]. In 1980, a conference in Villach, Austria, was hosted by the WMO, UNEP and the International Council of Scientific Unions (ICSU), with the aim of providing a "carefully prepared scientific assessment of the CO2 question to provide them with guidance in their future activities and advice to nations". This conference concluded that the scientific uncertainties were so significant that no CO2 management plan could be proposed. The same three organizations tried again in Villach in1985, using essentially the same data, but this time the 100 attendees participated as individuals rather than representatives of their countries, and they were selected by the three sponsoring agencies because of their support for global warming. This conference included the presentation of several papers, which were both commissioned and peer-reviewed by the conference organizers [Boehmer-Christiansen and Kellow, 2002]. Bert Bolin wrote the report for this conference and created a consensus on the need to take action on global warming. Bolin was also instrumental in preparing the SCOPE 29 report on ‘The greenhouse effect, climate change and ecosystems’ [Bolin et al, 1986]. This created the necessary political pressure for the WMO to establish the IPCC in 1988. Hansen also presented his fraudulent climate warming data to a US Congressional hearing in June 1988 [Hamlin, 2021].
When the IPCC was created in 1988, Bolin was the first chairman and another global warming believer, John Houghton, Director General of the UK Met Office led Working Group 1 for the technical assessment of global warming. The UK Hadley Center for Climate Prediction was established at the Met. Office in 1989. In conjunction with the Climate Research Center at the University of E. Anglia, the Hadley Center provided major support to the IPCC. The first IPCC assessment report was published in 1990. It was based largely on the SCOPE 29 report.
It must be emphasized that the Intergovernmental Panel on Climate Change (IPCC) is a political body, not a scientific one [McLean, 2010, 2009, Bolin, 2007]. Its mission is to assess “the scientific, technical and socioeconomic information relevant for the understanding of the risk of human-induced climate change.” This is based on the a-priori assumption that human activities are causing CO2 induced global warming. There never was an attempt to objectively evaluate the scientific evidence of the cause of climate change. The IPCC was established to exploit global warming as a way of inducing economic disruption based on the population control and sustainability concerns raised by the Club of Rome [Darwall, 2017, Zubrin, 2013, Klaus, 2007, Dewar 1995]. The IPCC has published five major assessment reports: the first, second and third - FAR (1990), SAR (1995), TAR (2001) and AR4 (2007), and AR5 (2013) and is in the process of publishing AR6 (2021). While the reports may contain a useful compendium of scientific references, material that does not conform to the global warming dogma has usually been omitted. Authors and editors were selected based on their willingness to find CO2 induced global warming whether it existed or not. The primary focus of these reports has been on the use of modeling ‘scenarios’ to predict future global warming using invalid computer models. These reports should not be cited as scientific references. Any scientific caution about the attribution of temperature increases to global warming was abandoned with the second IPCC Assessment Report in 1995. This was altered at the last minute at the request of the US State Department [FM, 2012]. The science had to agree with the ‘Summary for Policymakers’ written for the politicians. Similarly, the notorious ‘Hockey Stick’ temperature series based on fraudulent tree ring data was featured prominently in the 2001 Assessment Report [Mann et al, 1998, 1999, Montford, 2010, Steyn, 2015, Wedgman et al, 2010]. This was an attempt to eliminate the Medieval Warm Period and the Maunder Minimum from the climate record. The fraud here was the deliberate manipulation of the measured data to create the desired outcome.
Close ties developed between political leaders and various leading climate researchers. In the UK this included John Houghton (UK Met Office), the Climate Research Unit (CRU) at UEA and Margaret Thatcher (UK Prime Minister) [Courtney, 2012]. In the US one of leading political advocates of climate change was Al Gore. He first heard of global warming as a student when he took a course from Roger Revelle. Gore was elected to Congress in 1976 and was US Vice president from 1992 to 2000. He was later responsible for ‘An Inconvenient Truth’. This was a largely fraudulent book on global warming that was also made into a film of the same name. International efforts led to the Kyoto protocol which was established in 1997 and ratified in 2005. Some of the Kyoto provisions ended in 2012. The US never ratified the Kyoto Protocol. The Paris ‘accord’ is latest ‘voluntary’ attempt at some kind of agreement. One of the reasons for the failure is because the ‘developed’ countries have refused to pay the large sums of money ($100 billion per year) requested by the ‘underdeveloped’ countries for ‘damage’ caused by ‘global warming’. In addition, a large number of coal fired electrical power plants are either under construction or being planned notably by China and India. US policy towards global warming has also changed significantly with the last three US presidents elections.
In November of 2009, and again in November 2011, a large archive of e-mails and other files from the Climate Research Unit of the University of East Anglia was released on the Internet. A third round was released in March 2013. This archive has revealed to many people outside of the close knit climate community that there had been an ongoing fraud for many years to promote the global warming agenda and prevent the publication of material that did not support the prevailing global warming dogma. The peer review process in climate science had collapsed and been replaced by blatant cronyism. Climate science had become detached from its foundation in physical science and degenerated into a quasi-religious cult. Belief in global warming was a prerequisite for funding in climate science. The release of this climate archive became known as ‘Climategate’. The information provided has been analyzed in detail by several authors [Monckton, 2009, Montford 2010, Mosher & Fuller, 2010].
7.0 CONCLUSIONS
The Sixth IPCC Climate Assessment (AR6) should be rejected outright because the report is based on the results from fraudulent climate models. The fundamental error is the underlying assumption of radiative forcing in an equilibrium average climate. There are no forcings, feedbacks or climate sensitivity to CO2 in the earth’s climate. The modern modeling fraud can be traced back to two papers, Manabe and Wetherald [1967] and Hansen et al [1981]. Between them, these papers contain seven fundamental scientific errors that provided the foundation for the multi-trillion dollar climate fraud we have today. The fraud has at least three different parts. First there is the technical fraud. This started in the nineteenth century with the equilibrium climate hypothesis and speculation that changes in the atmospheric concentration of CO2 could cause the earth to cycle through an Ice Age. Later, exaggerated concern over fossil fuel combustion led to melodramatic prophecies of the global warming apocalypse. These became such a good source of research funding that the scientific process of hypothesis and discovery collapsed. Second, there was institutional fraud related to ‘mission creep’ within various government agencies. As funding for NASA and Atomic Energy decreased, government scientists found alternative employment as climate modelers. Third, there was a deliberated decision by various outside interests, including environmentalists and politicians to exploit the climate apocalypse to further their own causes. There was no single person or event that created the climate fraud. There was a gradual transition from the invalid hypothesis of an equilibrium average climate to the massive multi-trillion dollar pyramid or Ponzi scheme that we have today.
Manabe and Wetherald chose to ignore the limitations imposed by the assumptions in their 1967 paper and continued to build a primitive global circulation model. This created global warming as a mathematical artifact by definition based on the oversimplified and averaged energy transfer processes incorporated into the model. At NASA, there was a complete failure to apply the principles of technology readiness levels to the climate models. Hansen’s 1981 paper fails at Level 1, basic principles observed and reported. Wind driven ocean evaporation, the 1940 climate temperature peak and switch from 'equilibrium' surface and air temperatures to weather station temperatures were ignored. Similarly, the 1979 Charney report ignored the evidence about the real cause of an Ice Age and the non-equilibrium climate revealed by satellite observations. The work of Lorenz on weather forecasting model instability was also ignored. No thermal engineering calculations of the change in surface temperature were performed as model validation. The non-equilibrium phase shifts between the peak solar flux and the surface temperature response were never considered.
The peer review process in climate science has collapsed and been replaced by blatant cronyism. The climate modelers have retreated inside a cocoon of lies where they discuss the pseudoscience of radiative forcings, feedbacks and climate sensitivities to a CO2 ‘doubling’. This is obvious in the discussion of the climate models presented in Chapter 7 of the AR6 WGp I report. No quantitative discussion of the climate energy transfer processes that link radiative forcing to the measured surface temperature is presented. How does an increase of 0.034 W m-2 per year in the downward LWIR flux to the surface cause any kind of climate change? How does this create ‘extreme weather events’? How does the LWIR component of a radiative forcing heat the ocean when it cannot penetrate below 100 micron into the ocean surface? How does the radiative forcing couple to the wind driven ocean evaporation? How much of the observed temperature change can be explained by the ocean oscillations? Where are the convection transition temperature and the solar flux-temperature phase shifts discussed? How are heating effects from downslope winds and blocking high pressure systems separated from those produced by ‘radiative forcing’?
The ‘advanced’ climate models are used to create the sacred spaghetti plots of global warming. This is GIGO: garbage in, gospel out. The climate modelers are simply playing expensive computer games in an equilibrium climate fantasy land. The model ‘predictions’ are fed directly to government agencies and the IPCC with minimal outside scrutiny. There is no climate science involved. Irrational belief in computer models has replaced the Laws of Physics. The Imperial Cult of the Global Warming Apocalypse has claimed the Divine Right to save the world from a non-existent problem. There is no ‘climate crisis’. Eisenhower’s warning about the corruption of science by government funding has come true. The entire multi-trillion dollar Ponzi or pyramid scheme built on these fraudulent modeling results needs to be shut down and those responsible should face the legal consequences of their activities.
ACKNOWLEDGEMENT
This work was performed as independent research by the author. It was not supported by any grant awards and none of the work was conducted as a part of employment duties for any employer. The views expressed are those of the author. He hopes that you will agree with them.
REFERENCES
Normally, the references given in an article of this nature would be almost exclusively to the peer reviewed literature, with limited references to websites that provide access to climate data. Unfortunately, climate science has been thoroughly corrupted by the global warming fraud. The peer review process has collapsed and been replaced by blatant cronyism. Many of the publications in ‘prestigious’ journals such as Nature, Science, PNAS and others that relate to climate modeling predictions of global warming are fraudulent and should never have been published. Consequently many of the important references given here are to website publications. This should not detract from the integrity of the information provided. Many of these website publications have received a more thorough review than they might have received through the traditional peer review process.
Agassiz, L., Etudes sur les Glaciers, Neuchatel, 1840
AMO, 2020 https://www.esrl.noaa.gov/psd/data/correlation/amon.us.long.mean.data Andrews, R., 2017a, Energy Matters Sept 14, 2017, ‘Adjusting Measurements to Match the Models – Part 3: Lower Troposphere Satellite Temperatures’. http://euanmearns.com/adjusting-measurements-to-match-the-models-part-3-lower-troposphere-satellite-temperatures/#more-19464
Andrews, R., 2017b, Energy Matters Aug 2, 2017, ‘Making the Measurements Match the Models – Part 2: Sea Surface Temperatures’. http://euanmearns.com/making-the-measurements-match-the-models-part-2-sea-surface-temperatures/
Andrews, R., 2017c, Energy Matters July 27, 2017, ‘Adjusting Measurements to Match the Models – Part 1: Surface Air Temperatures’. http://euanmearns.com/adjusting-measurements-to-match-the-models-part-1-surface-air-temperatures/
Argo, 2020, https://argo.ucsd.edu/data/data-visualizations/marine-atlas/
Arrhenius, S., Philos. Trans. 41 237-276 (1896), ‘On the influence of carbonic acid in the air upon the temperature of the ground’. http://empslocal.ex.ac.uk/people/staff/gv219/classics.d/Arrhenius96.pdf
Billo, E. J., Excel for Scientists and Engineers, J. Wiley & Sons, Hoboken, NJ, USA, 2007 https://www.academia.edu/37366688/Joseph_Billo_Excel_for_Scientists_and_Engineers_Numerical_Methods_pdf
Boehmer-Christiansen, S. and A.J. Kellow, ‘International environmental policy: interests and the failure of the Kyoto process’, Northampton, MA, Edward Elgar Pub., 2002.
Bolin, B., ‘A History of the Science and Politics of Climate Change. The Role of the Intergovernmental Panel on Climate Change’, Cambridge, Cambridge University Press, 2007
Bolin, B., Tellus 12 274-281 (1960), ‘On the Exchange of Carbon Dioxide between the Atmosphere and the Sea’, http://onlinelibrary.wiley.com/doi/10.1111/j.2153-3490.1960.tb01311.x/epdf
Bolin, B., B, Döös, J. Jäger and R.A Warwick, The Greenhouse Effect, Climate Change and Ecosystems, SCOPE 29, John Wiley and Sons, 1986,
Bolin, B. and E. Eriksson, ‘Changes in the Carbon Dioxide Content of the Atmosphere and Sea Due to Fossil Fuel Combustion’, in ‘The Atmosphere and the Sea in Motion’, edited by Bert Bolin, pp. 130-42. New York: Rockefeller Institute Press, 1959,
Bolin, B. and E. Eriksson, ‘Distribution of matter in the atmosphere and the sea in motion’, in Changes in the Carbon Dioxide Content of the Atmosphere and Sea Due to Fossil Fuel Combustion, edited by Bert Bolin, pp. 130-42. New York: Rockefeller Institute Press, 1959. https://math.nyu.edu/~gerber/courses/2018-fruhling/bolin_etal-atmoshere_sea_in_motion-1959.pdf
BP, 2020, ‘Statistical Review of World Energy’ 2020, https://www.bp.com/en/global/corporate/energy-economics/statistical-review-of-world-energy.html
Bryson, R. A. and G. J. Dittberner, J. Atmos. Sci. 33(11) 2094-2106 (1976), ‘A non-equilibrium model of hemispheric mean surface temperature’. https://journals.ametsoc.org/view/journals/atsc/33/11/1520-0469_1976_033_2094_anemoh_2_0_co_2.xml
Callendar, G. S., J. Roy. Met. Soc. 64 223-240 (1938), ‘The artificial production of carbon dioxide and its influence on temperature’, https://doi.org/10.1002/qj.49706427503, availalbe at: http://www.met.reading.ac.uk/~ed/callendar_1938.pdf
CERES 2011, CERES OLR Image, March 8 2011, Aqua Mission (EOS/PM-1), https://earth.esa.int/web/eoportal/satellite-missions/a/aqua
Charney, J. G. et al, Carbon Dioxide and Climate: A Scientific Assessment, Report of an ad hoc study group on carbon dioxide and climate, Woods Hole, MA July 23-27 (1979), https://www.bnl.gov/envsci/schwartz/charney_report1979.pdf
Clark, R., 2013, Energy and Environment 24(3, 4) 319-340 (2013), ‘A dynamic coupled thermal reservoir approach to atmospheric energy transfer Part I: Concepts’. https://doi.org/10.1260/0958-305X.24.3-4.319
Energy and Environment 24(3, 4) 341-359 (2013), ‘A dynamic coupled thermal reservoir approach to atmospheric energy transfer Part II: Applications’. https://doi.org/10.1260/0958-305X.24.3-4.341
Courtney, R. 2012, ‘The History of the Global Warming Scare’ https://tallbloke.wordpress.com/2012/09/12/richard-courtney-the-history-of-the-global-warming-scare/
Darwall, R., ‘Green Tyranny’ Encounter Books, NY, NY, 2017
Dewar, E., 'Cloak of Green: The Links between Key Environmental Groups, Government and Big Business', Lorimer Press, 1995
Douglas, J. H., Science News 107 138-140 March 1, (1975), ‘Climate change: chilling possibilities’, https://www.sciencenews.org/wp-content/uploads/2008/10/8983.pdf
Ehrlich, P. R., The Population Bomb, Ballantine Books, New York, 1968, http://pinguet.free.fr/ehrlich68.pdf
ENSO, 2020, https://psl.noaa.gov/gcos_wgsp/Timeseries/Data/nino34.long.data
FM, 2012, http://fabiusmaximus.com/2012/09/11/ipcc-climate-change-science-warming-42859/#more-42859
Fourier, B. J. B., Mem. R. Sci. Inst., 7 527-604 (1827), ‘Memoire sur les temperatures du globe terrestre et des espaces planetaires’. https://www.academie-sciences.fr/pdf/dossiers/Fourier/Fourier_pdf/Mem1827_p569_604.pdf
English translation: https://courses.seas.harvard.edu/climate/eli/Courses/EPS281r/Sources/Greenhouse-effect/more/Fourier-1827.pdf
Fourier, B. J. B., Annales de Chimie et de Physique, 27, pp. 136–167 (1824), ‘Remarques générales sur les températures du globe terrestre et des espaces planétaires’. https://gallica.bnf.fr/ark:/12148/bpt6k65708960/f142.image#
English translation:
http://fourier1824.geologist-1011.mobi/
Fourier, B. J. B., Theorie analytique de la chaleur, Didot, Paris, 1822 https://gallica.bnf.fr/ark:/12148/bpt6k29061r/f7.item
Garai, A. and J. Kleissl, J., Atmos. Sci. 68(12) 2945-2954 (2011), ‘Air and Surface Temperature Coupling in the Convective Atmospheric Boundary Layer’, https://doi.org/10.1175/JAS-D-11-057.1
Gerlich, G. and R. D. Tscheuschner, Int. J. Mod. Phys. B, 23(3) 274-394 (2009) ‘Falsification of the atmospheric CO2 greenhouse effects within the frame of physics’, https://doi.org/10.1142/S021797920904984X
Gray, S. T., L. J. Graumlich, J. L. Betancourt and G. T. Pederson, Geophys. Res. Letts, 31 L12205, pp1-4 (2004) doi:10.1029/2004GL019932, ‘A tree-ring based reconstruction of the Atlantic Multi-decadal Oscillation since 1567 A.D.’.http://www.riversimulator.org/Resources/ClimateDocs/GrayAMO2004.pdf,
Gray, S. T., et al. 2004, Atlantic Multi-decadal Oscillation (AMO) Index Reconstruction, IGBP PAGES/World Data, Center for Paleoclimatology, Data Contribution Series #2004-062, NOAA/NGDC Paleoclimatology Program, Boulder CO, USA. https://www.ncei.noaa.gov/pub/data/paleo/treering/reconstructions/amo-gray2004.txt
HadCRUT4, 2019, https://www.metoffice.gov.uk/hadobs/hadcrut4/data/current/time_series/HadCRUT.4.6.0.0.annual_ns_avg.txt
Hale, G. M. and Querry, M. R., Applied Optics, 12(3) 555-563 (1973), ‘Optical constants of water in the 200 nm to 200 µm region’. https://doi.org/10.1364/AO.12.000555
Hamlin, L., WUWT Post 4/22/21, ‘History Confirms Democrat’s 1988 Senate Global Warming Hearing Got Everything Wrong from Start to Finish’, https://wattsupwiththat.com/2021/04/22/history-confirms-democrats-1988-senate-global-warming-hearing-got-everything-wrong-from-start-to-finish/
Hanel, R. A., B. Schlachman, D. Rogers and D. Vanous, Applied Optics 10(6) 1376-1382 (1971), ‘Nimbus 4 Michelson Interferometer’, https://doi.org/10.1364/AO.10.001376
Hansen, J. et al., (45 authors), J. Geophys Research 110 D18104 pp1-45 (2005), ‘Efficacy of climate forcings’. https://pubs.giss.nasa.gov/docs/2005/2005_Hansen_ha01110v.pdf
Hansen, J., D. Johnson, A. Lacis, S. Lebedeff, P. Lee, D. Rind and G. Russell Science 213 957-956 (1981), ‘Climate impact of increasing carbon dioxide’. https://pubs.giss.nasa.gov/docs/1981/1981_Hansen_ha04600x.pdf
Harde, H., Int. J. Atmos. Sci.9251034 (2017), ‘Radiation Transfer Calculations and Assessment of Global Warming by CO2’. https://doi.org/10.1155/2017/9251034
Hausfather, Z., ‘CMIP6: The next generation of climate models explained’ Carbon Brief, 2019, https://www.carbonbrief.org/cmip6-the-next-generation-of-climate-models-explained Hays, J. D., J. Imbrie, N. J. Shackleton, Science, 194 Dec. 10, pp 1121-1132 (1976), ‘Variations in the Earth's Orbit: Pacemaker of the Ice Ages’, https://www.science.org/doi/10.1126/science.194.4270.1121
Hecht, M. M., 21st Century Science and Technology, pp.64-68, Fall 2007, ‘Where the global warming hoax was born’ http://www.21stcenturysciencetech.com/Articles%202007/GWHoaxBorn.pdf
Iacono, M. J., J. S. Delamere, E. J. Mlawer, M. W. Shephard, S. A. Clough, and W. D. Collins, J. Geophys. Res., 113, D13103pp 1-8, (2008), ‘Radiative forcing by long-lived greenhouse gases: Calculations with the AER radiative transfer models’. https://doi.org/10.1029/2008JD009944
Imbrie, J. and K. P. Imbrie, ‘Ice Ages: Solving the Mystery’, Harvard University Press, Cambridge, Mass, 1979
IOD, 2020, https://psl.noaa.gov/gcos_wgsp/Timeseries/Data/dmi.had.long.data
IPCC, 2021: Climate Change 2021: The Physical Science Basis. Contribution of Working Group I to the Sixth Assessment Report of the Intergovernmental Panel on Climate Change [Masson-Delmotte, V., P. Zhai, A. Pirani, S.L. Connors, C. Péan, S. Berger, N. Caud, Y. Chen, L. Goldfarb, M.I. Gomis, M. Huang, K. Leitzell, E. Lonnoy, J.B.R. Matthews, T.K. Maycock, T. Waterfield, O. Yelekçi, R. Yu, and B. Zhou (eds.)]. Cambridge University Press. In Press. https://www.ipcc.ch/report/sixth-assessment-report-working-group-i/
IPCC, 2013: Myhre, G., D. Shindell, F.-M. Bréon, W. Collins, J. Fuglestvedt, J. Huang, D. Koch, J.-F. Lamarque, D. Lee, B. Mendoza, T. Nakajima, A. Robock, G. Stephens, T. Takemura and H. Zhang, ‘Anthropogenic and Natural Radiative Forcing’. In: Climate Change 2013: The Physical Science Basis. Contribution of Working Group I to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change [Stocker, T.F., D. Qin, G-K. Plattner, M. Tignor, S.K. Allen, J. Boschung, A. Nauels, Y. Xia, V. Bex and P.M. Midgley (eds.)]. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA, Chapter 8, Radiative Forcing1535 pp, doi:10.1017/CBO9781107415324. http://www.climatechange2013.org/report/full-report/
Jones, P. D., T. M. Wigley and P. B Wright, Nature 323 (31) 430-434 (1986), ‘Global temperature variations between 1861 and 1984’. https://www.nature.com/articles/322430a0 Available at: https://geosci.uchicago.edu/~archer/warming_papers/jones.1986.temp_record.pdf
Keeling, 2021, https://scripps.ucsd.edu/programs/keelingcurve/
Klaus, V., Blue Planet in Green Shackles. What Is Endangered: Climate or Freedom? Competitive Enterprise Institute, 2007
Kluft, L., Reports on Earth System Science / Max Planck Institute for Meteorology 239 (2020), ‘Benchmark Calculations of the Climate Sensitivity of Radiative-Convective Equilibrium’. https://pure.mpg.de/rest/items/item_3274272/component/file_3274608/content
Knutti, R. and G. C. Hegerl, Nature Geoscience 1 735-743 (2008), ‘The equilibrium sensitivity of the Earth’s temperature to radiation changes’. https://www.nature.com/articles/ngeo337
Koll, D. D. B and T. W. Cronin., PNAS, (2018), ‘Earth’s outgoing longwave radiation linear due to H2O greenhouse effect’. https://www.pnas.org/content/115/41/10293
Lacis, A. A. and V. Oindas, J. Geographical Res. 96(D5) 9027-9063 (1991), ‘A description of the correlated k distributing method for modeling non gray gaseous absorption, thermal emission and multiple scattering in vertically inhomogeneous atmospheres’, https://doi.org/10.1029/90JD01945
Lamb, H. H., ‘Climate, Past, Present and Future’, V1, Routledge, Abingdon, UK. 1972 Lettau, H.H., and B. Davidson , ‘Exploring the atmosphere’s first mile’, Oxford: Pergamon Press, 1957, https://doi.org/10.1002/qj.49708436022
Lewis, N. and J. Curry, J. Climate 31 6051-6070 (2018), ‘The Impact of Recent Forcing and Ocean Heat Uptake Data on Estimates of Climate Sensitivity’. https://journals.ametsoc.org/view/journals/clim/31/15/jcli-d-17-0667.1.xml
Lorenz, E. N., J. Applied Meteorology and Climatology 12(3) 543-546 (1973), ‘On the Existence of Extended Range Predictability’, https://journals.ametsoc.org/view/journals/apme/12/3/1520-0450_1973_012_0543_oteoer_2_0_co_2.xml?tab_body=fulltext-display Lorenz, E.N., Journal of the Atmospheric Sciences 20, pp. 130-41 (1963), ‘Deterministic non-periodic flow.’ https://journals.ametsoc.org/view/journals/atsc/20/2/1520-0469_1963_020_0130_dnf_2_0_co_2.xml
Luning, S. and F. Vahrenholt, Frontiers in Earth Science 5:104 pp 1-7 (Dec. 2017) ‘Paleoclimatological context and reference level of the 2 C and 1.5 C Paris agreement long-term temperature limits’, https://www.frontiersin.org/articles/10.3389/feart.2017.00104/full Manabe S. and F. Moller, Monthly Weather Review 89(12) 503-532 (Dec. 1961), ‘On the radiative equilibrium and heat balance of the atmosphere’, https://journals.ametsoc.org/view/journals/mwre/89/12/1520-0493_1961_089_0503_otreah_2_0_co_2.xml
Manabe, S. and R. T. Wetherald, J. Atmos. Sci. 32(1) 3-15 (1975), ‘The effects of doubling the CO2 concentration in the climate of a general circulation model’. https://journals.ametsoc.org/view/journals/atsc/32/1/1520-0469_1975_032_0003_teodtc_2_0_co_2.xml?tab_body=pdf
Manabe, S. and R. T. Wetherald, J. Atmos. Sci., 24 241-249 (1967), ‘Thermal equilibrium of the atmosphere with a given distribution of relative humidity’. http://www.gfdl.noaa.gov/bibliography/related_files/sm6701.pdf
Mann M. E., R S. Bradley and M. K. Hughes, Geophys Res Lett 26:759–762 (1999), ‘Northern Hemisphere temperatures during the past millennium: Inferences, uncertainties, and limitations’. http://www.meteo.psu.edu/holocene/public_html/shared/articles/MBH1999.pdf
Mann, M. E.: R. E. Bradley and M. K. Hughes, Nature 392, 779-787 (1998) ‘Global-scale temperature patterns and climate forcing over the past six centuries’ http://www.meteo.psu.edu/holocene/public_html/shared/articles/mbh98.pdf
McFarlane, F., Watts Up with That, 11.19.2018, ‘The 1970s Global Cooling Consensus was not a Myth’, https://wattsupwiththat.com/2018/11/19/the-1970s-global-cooling-consensus-was-not-a-myth/
McLean, J., 2010, ‘we have been conned – an independent review of the IPCC’, SPPI 2010 http://scienceandpublicpolicy.org/originals/we_have_been_conned.html
McLean, J., 2009, ‘Climate Science Corrupted’, SSPI 2010, http://scienceandpublicpolicy.org/images/stories/papers/originals/climate_science_corrupted.pdf
Mead, M. and W. W. Kellogg eds, The Atmosphere: Endangered and Endangering, Fogarty International Center Proceedings No. 39, 1976 (Washington, D.C.: U.S. Government PrintingOffice, DHEW Publication No. [NIH] 77-1065). (Google Digital Book)
Meadows, D. H., D. L. Meadows, J. Randers and W. W. Behrens, ‘Limits to Growth’, Universe Books, NY, 1972,
Meehl, G. A., G. J. Boer, C. Covey, M. Latif and R. J. Stouffer, Eos, 78(41) pp445-451 October 14 (1997), ‘Intercomparison Makes for a Better Climate Model’ https://doi.org/10.1029/97EO00276
Meinshausen, M, S. J. Smith, K. Calvin, J. S. Daniel, M. L. T. Kainuma, J-F. Lamarque, K. Matsumoto, S. A. Montzka, S. C. B. Raper, K. Riahi, A. Thomson, G. J. M. Velders and D.P. P. van Vuuren, Climatic Change 109:213–241 (2011) DOI 10.1007/s10584-011-0156-z, ‘The RCP greenhouse gas concentrations and their extensions from 1765 to 2300’. https://link.springer.com/content/pdf/10.1007/s10584-011-0156-z.pdf
MODTRAN, 2020 http://forecast.uchicago.edu/Projects/modtran.orig.html
Moller, F., Applied Optics 3(2) 157-166 (1964), ‘Optics of the lower atmosphere’, https://doi.org/10.1364/AO.3.000157
Monckton, C., SPPI, 2009: ‘Climategate: caught green-handed’, http://scienceandpublicpolicy.org/monckton/climategate.html
Montford, A. W., ‘The Hockey Stick Illusion’, Stacey International, 2010
Mosher, S. and T. W. Fuller, 'Climategate: The Crutape Letters', Create Space, 2010.
Oke T. R., WMO/TD-No. 1250, World Meteorological Association, 2006, ‘Initial guidance to obtain representative meteorological observations at urban sites’. https://www.researchgate.net/publication/265347633_Initial_guidance_to_obtain_representative_meteorological_
Otto, A., F. E. L. Otto, O. Boucher, J. Church, G. Hegerl, P. M. Forster, N. P. Gillett, J. Gregory, G. C. Johnson, R Knutti, N. Lewis, U. Lohmann, J. Marotzke, G. Myhre, D. Shindell, B. Stevens and M. R. Allen, Nature Geoscience, 6 (6). 415 - 416 (2013). ISSN 1752-0894, ‘Energy budget constraints on climate response’. http://eprints.whiterose.ac.uk/76064/7/ngeo1836(1)_with_coversheet.pdf
Otto, A., F. E. L. Otto, O. Boucher, J. Church, G. Hegerl, P. M. Forster, N. P. Gillett, J. Gregory, G. C. Johnson, R Knutti, N. Lewis, U. Lohmann, J. Marotzke, G. Myhre, D. Shindell, B. Stevens and M. R. Allen, Nature Geoscience, 6 (6). 415 - 416 (2013). ISSN 1752-0894, ‘Energy budget constraints on climate response’, Supplementary Material. https://static-content.springer.com/esm/art%3A10.1038%2Fngeo1836/MediaObjects/41561_2013_BFngeo1836_MOESM299_ESM.pdf
PDO, 2020, https://www.ncdc.noaa.gov/teleconnections/pdo/
Peterson, T. C., W. M. Connolley and J. Fleck, Bull. Amer. Meteor. Soc., 86 1325-1337 (2008), ‘The myth of the 1970’s global cooling consensus’ https://doi.org/10.1175/2008BAMS2370.1
Plass, G.N., Tellus 8(2) 140-154 (1956), ‘The Carbon Dioxide Theory of Climatic Change’, https://onlinelibrary.wiley.com/doi/abs/10.1111/j.2153-3490.1956.tb01206.x
Plass, G.N. , Quarterly Journal of the Royal Meteorological Society 82: 310-324 (1956), ‘The influence of the 15-micron carbon dioxide band on the atmospheric infrared cooling rate’, https://doi.org/10.1002/qj.49708235307
available at: http://www.rescuethatfrog.com/wp-content/uploads/2017/02/Plass-1956c.pdf Pouillet, M., in: Scientific Memoirs selected from the Transactions of Foreign Academies of Science and Learned Societies, edited by Richard Taylor, 4 (1837), pp. 44-90. ‘Memoir on the solar heat, on the radiating and absorbing powers of the atmospheric air and on the temperature of space’ http://nsdl.library.cornell.edu/websites/wiki/index.php/PALE_ClassicArticles/archives/classic_articles/issue1_global_warming/n2-Poulliet_1837corrected.pdf
Original publication: Comptes Rendus des Séances de l'Académie des Sciences. Paris. 7, 24-65 (1836). ‘Mémoire sur la chaleur solaire : sur les pouvoirs rayonnants et absorbants de l'air atmosphérique et sur la température de l'espace’. https://gallica.bnf.fr/ark:/12148/bpt6k95017r.image
Poyet, P. The Rational Climate e Book, Malta, 2020 e-ISBN 978-99957-1-929-6, https://www.researchgate.net/publication/347150306_The_Rational_Climate_e-Book/link/5fe21ddb92851c13feb1763d/download
Ramanathan, V., Science, 190, 50–52, (1975), ‘Greenhouse effect due to chlorofluorocarbons: Climatic implications’, https://www.science.org/doi/abs/10.1126/science.190.4209.50
Ramanathan, V. and J. A. Coakley, Rev. Geophysics and Space Physics 16(4)465-489 (1978), ‘Climate modeling through radiative convective models’. https://doi.org/10.1029/RG016i004p00465 ,
Revelle, R. and H. E. Suess, Tellus 9 18-27 (1957), ‘Carbon Dioxide Exchange between Atmosphere and Ocean and the Question of an Increase of Atmospheric CO2 During the Past Decades’, https://pdfs.semanticscholar.org/d014/06a57bff758203390e36247bd96e0c9f8102.pdf
Spencer, R., 2021, https://www.drroyspencer.com/2021/04/an-earth-day-reminder-global-warming-is-only-50-of-what-models-predict/
Steyn, M., “A Disgrace to the Profession”, Amazon, 2015, https://www.amazon.com/%2522A-Disgrace-Profession%2522-Steyn-editor/dp/0986398330/ref=sr_1_2?s=books&ie=UTF8&tag=wattsupwithth-20&qid=1439321074&sr=1-2&keywords=%22A+Disgrace+to+the+Profession%22
Stouffer, R. J., V. Eyring, G. A. Meehl, S. Bony, C. Senior, B. Steven, S, and K. E. Taylor, BAMS 98(1) 95-105 (2017), ‘CMIP5 scientific gaps and recommendations for CMIP6’ https://journals.ametsoc.org/doi/pdf/10.1175/BAMS-D-15-00013.1
Taylor, F. W., Elementary Climate Physics, Oxford University Press, Oxford, 2006, Chapter 7.
Taylor, K. E., R. J. Stauffer and G. A. Meehl, Bull. Amer. Met. Soc. 93(4) 485-498 (2012), ‘An overview of the CMIP5 and the experimental design’ https://doi.org/10.1175/BAMS-D-11-00094.1
Tyndall, J., Philosophical Transactions of the Royal Society of London, 151 pp. 1-36, (1861), ‘On the Absorption and Radiation of Heat by Gases and Vapours, and on the Physical Connexion of Radiation, Absorption, and Conduction’. https://www.jstor.org/stable/pdf/108724.pdf?refreqid=excelsior%3A86f44613289e621f0b8d37d54020bf60
Tyndall, J., Proc. Roy Inst. Jan 23 pp 200-206 (1863), ‘On radiation through the Earth's atmosphere’.
Wegman, E. J., D. W. Scott & Y. H. Said, 2010, ‘Ad hoc committee report on the 'hockey stick' global climate reconstruction’ http://scienceandpublicpolicy.org/reprint/ad_hoc_report.html
Wijngaarden, W. A. van and W. Happer, arXiv:2006.03098v1 [physics.ao-ph] 4 Jun 2020, ‘Dependence of earth's thermal radiation on five most abundant greenhouse gases’. https://arxiv.org/abs/2006.03098
Wilson, R. C., Astrophys Space Sci DOI 10.1007/s10509-014-1961-4, pp 1-14 (2014), ‘ACRIM3 and the Total Solar Irradiance database’ https://www.researchgate.net/publication/262376602_ACRIM3_and_the_Total_Solar_Irradiance_database/download
WRCC, 2020, https://wrcc.dri.edu/sod/arch/hbF.html
Yu, L., Jin, X. and Weller R. A., OAFlux Project Technical Report (OA-2008-01) Jan 2008, ‘Multidecade Global Flux Datasets from the Objectively Analyzed Air-sea Fluxes (OAFlux) Project: Latent and Sensible Heat Fluxes, Ocean Evaporation, and Related Surface Meteorological Variables’. https://rda.ucar.edu/datasets/ds260.1/docs/OAFlux_TechReport_3rd_release.pdf
Zubrin, R., ‘Merchants of Despair’, Encounter Books, NY, NY, 2013