This is an HTML version of an attachment to the Official Information request 'Records mentioning scientists Dr Ronan Connolly and/or Michael Connolly and/or Christopher Monckton'.

ACT 1982
INFORMATION 
RELEASED UNDER THE 
 OFFICIAL 

149 
Hermann Harde: What Humans Contribute to Atmospheric CO2: Comparison of Carbon Cycle Models with Observations 
considers  mean  values  of  the  net  atmospheric  accumulation 
reservoirs  on  atmospheric  CO2,  details  of  other  extraneous 
<dC/dt> = 1.7 ppm/yr and of the human emissions <dCA/dt> =
reservoirs of carbon are entirely irrelevant. This feature of the 
eA(t) = 3 ppm/yr in a balance
governing physics is not only powerful, but fortunate.  
Concerning carbonate chemistry, it is noteworthy that, in 
dC dt − dC dt

 (26) 
A
dC dt
N
< 0
the Earth’s distant past, CO2 is thought to have been almost 
2000% as great as its present concentration (e.g., Royer et. 
in  which  with  <dCA/dt>  =  eA(t)  a  priori  any  anthropogenic
al. [30]). Most of that was absorbed by the oceans, in which 
absorptions  are  embezzled.  From  this  relation  it  is  also 
carbon  today  vastly  exceeds  that  in  the  atmosphere. 
inferred  that  the  average  natural  contribution  <dCN/dt>  has 
According  to  the  IPCC,  even  in  modern  times  the  oceans 
been to remove CO2 from the atmosphere, this with the same 
account  for  40%  of  overall  absorption  of  CO2  (AR5  [1], 
wrong conclusion as Cawley that the long term trend of rising 
Fig.6.1). In relation to other sinks, their absorption of CO2 
CO2 could not be explained by natural causes. This argument 
is clearly not limited (see Appendix A). Of that 40%, over 
is  disproved  with  Figures  8  and  10.  The  fact  that  the 
the  Industrial  Era  anthropogenic  CO2  represents  less  than 
environment has acted as a net sink throughout the Industrial 
1%. Contrasting with that minor perturbation in absorption 
Era is a consequence of a dynamic absorption rate, which is 
is  oceanic  emission  of  CO2.  Through  upwelling  of 
only controlled by the total CO2 concentration C = CN + CA
carbon-enriched  water,  the  oceans  significantly  enhance 
So, also with additional native emissions and/or temperature 
natural emission of CO  (Zhang [31]).  
changes in the absorptivity the total uptake always tries - with 
Different  to  our  approach,  which  takes  into  account 
some time delay - to compensate for the total emissions which, 
human  and  also  naturally  varying  emissions  and 
of  course,  also  include  the  anthropogenic  fraction.  In  other 
absorptions, the models in Section 3 emanate from such a 
words:  Since  nature  cannot  distinguish  between  native  and 
simple  and  apparently  flawed  description  that  over 
human emissions, nature is always a net sink as long as human 
thousands of years CO2 was circulating like an inert gas in a 
ACT 1982
emissions  are  not  zero.  Thus,  except  for  shorter  temporary 
closed system, and only with the industrial revolution this 
events like volcanic activities the environment will generally 
closed cy le came out of control due to the small injections 
act  as  a  net  sink  even  in  the  presence  of  increasing  natural 
by human emi sion .  
emissions.  
To  equate  <dCA/dt>  in  (26)  exclusively  with  human
5.5. Different Time Constants 
emissions violates conservation of mass. Only when replacing 
<dC
The  different  time  scales  introduced  with  the  models  in 
A/dt> by <eA(t) - CA/τR>, eq.(26) satisfies the Conservation
Law,  and  when  additionally  replacing  <dC
Section  3  represent  different  absorption  processes  for  the 
N/dt>  by  <eN(t)  -
C
uptake  of  atmospheric  CO2  molecules  by  the  extraneous 
N/τR> eq.(26) converts to (23).
Again we emphasize that a separate treatment of the native 
reservoirs.  From  physical  principles  it  is  impossible  that  an 
and human cycle with their respective concentrations C
absorption process would differentiate between naturally and 
A  and 
C
anthropogenically 
emitted 
molecules. 
The 
temporal 
N is possible if and only if no contributions are missing and 
the two balances are linked together in on  rate equ tion with 
absorption  or  sequestration  -  except  for  smallest  corrections 
only one unitary residence time.  
due to isotopic effects - is for all molecules identical.  
The  absorption  also  cannot  decline  unexpectedly  by  more 
5.4. Too Simple Model 
than one order of magnitude with the begin of the Industrial 
INFORMATION 
Era  or  because  of  an  additional  emission  rate  of  a  few  %. 
Often  climate  scientists  argue  that  ch nges  of  CO2  in  the 
Observations  show  that  no  noticeable  saturation  over  recent 
atmosphere  cannot  be  understood  without  considering 
years could be found (Appendix A). 
RELEASED UNDER THE 
changes  in  extraneous  systems  (see  e.g.,  AR5  [1],  Chap.6; 
Oceans  and  continents  consist  of  an  endless  number  of 
Köhler et al. [8]), and they characterize the Conservation Law 
sources  and  sinks  for  CO2  which  act  parallel,  emitting  CO2 
as  a  flawed  1-box  description    because,  a  single  balance 
into the atmosphere and also absorbing it again. In the same 
equation would not account for details in other reservoirs. In 
way as the different emission rates add up to a total emission, 
particular,  they  refer  to  carbonate  chemistry  in  the  ocean, 
the  absorption  rates  with  individual  absorptivities  αi  -  and 
where CO2 is  mostly converted to bicarbonate ions. As only 
each  of  them  scaling  proportional  to  the  actual  CO2 
about  1%  rema ns  in  the  form  of  dissolved  CO2,  they  argue 
concentration - add up to a total uptake as a collective effect 
that  only  this  small  fraction  could  be  exchanged  with  the 
 OFFICIAL 
atmosphere.  Due  to  this  so-called  Revelle  effect,  carbonate 
= α C
1
+α C
2
+ ... +α C
T
N

 (27) 
chemistry 
would  sharply  limit  oceanic 
uptake 
of 
= α
( 1 + α2 + ... + α ) ⋅= α ⋅C
N
R
anthropogenic CO2. 
In  regard  to  understanding  changes  of  CO2  in  the 
Collective  absorption  thus  leads  to  exponential  decay  of 
atmosphere,  changes  in  extraneous  systems  are  only 
perturbation CO2 at a single rate  
qualifiedly  of  interest.  The  governing  law  of  CO2  in  the 
atmosphere  (4)  and  in  more  elaborate  form  (23)  is  self 
α = 1/τ = α

 (28) 
1 + α 2 + ... + α
R
R
N
contained.  With  the  inclusion  of  the  surface  fluxes  eT(t)  and 
a
This decay rate is faster than the rate of any individual sink 
T(t)  =  C  R(t),  which  account  for  influences  of  the  adjacent

Earth Sciences 2019; 8(3): 139-159 
150 
and it prevails as long as its concentration or its difference to 
as  the  main  drivers  for  the  observed  CO2  increase  in  the 
external  reservoirs  remains  nonzero  (see:  Harde  [6];  Salby 
atmosphere and also for the continuous climate changes over 
[11]).  
the past and present times.  
The above behavior is a consequence of the Conservation 
The  various  mechanisms,  along  with  their  dependence  on 
Law and in contrast to the Bern Model, where decay proceeds 
temperature  and  other  environmental  properties,  could  not 
at  multiple  rates.  A  treatment  of  CO2  with  a  multiple 
have  remained  constant  during  the  pre-industrial  era.  This 
exponential decay obeys the following: 
inconsistency  invalidates  the  fundamental  assumption,  that 
natural  emission  and  absorption  during  the  pre-industrial 
−α t
α
α
1
− t
2

t
C e
10
C e
20
+ ... + C e N
0

 (29) 
period  did  remain  constant.  Even  less  this  is  valid  over  the 
C
Industrial Era, a period which is characterized by the IPCC as 
1 + C2 + ... + CN
the fastest rise in temperature over the Holocene or even the 
Then differentiation gives: 
last interglacial.  
So,  the  CO
dC
2  partial  pressure  in  sea  water  approximately 
= −α C e−α t
α
α
1
2
α
α
changes  with  temperature  as  (pCO
1
10
− C e− t...
2
20

C e− t
N
2)sw(T)  = 
pCO2) w(T0)* 
dt
N
0
exp[0.0433*(T-T
= −α
 (30) 
0)]  (see:  Takahashi  et  al.  [32])  and  thus,  an 
C
α
α
1
1 −
...
2
2

C
N
N
increase  of  1°C  causes  a  pressure  change  of  about  18  µatm
≠ −(α α
α
1 +
2 + ... +
) ⋅ C
N
which  amplifies  the  influx  and  attenuates  the  outflux.  From 
observations over the North Atlantic Ocean (see, Benson et al. 
At multiple decay rates the corresponding sinks operate, not 
[33])  it  can  be  estimated  that  a  pressure  difference  ∆pCO2
collectively, but independently. After a couple of their decay 
between the atmosphere and ocean of 1 µatm contributes to a 
times,  the  fastest  sinks  become  dormant.  Overall  decay  then 
flux change of δfin ≈ 0.075 mol/m2/yr = 3.3 g/m2/yr. Therefore,
continues  only  via  the  slowest  sinks,  which  remove  CO2 
with an Earth s surface of 320 Mio. km2 covered by oceans and 
ACT 1982
gradually.  It  is  for  this  reason  that  such  a  treatment  leaves 
a  pressure  change  of  ∆pCO2 = 18 µatm, under conventional
atmospheric  CO2 perturbed for longer than a thousand  years 
conditions  the  native  influx  from  oceans  to  the  atmosphere 
(Figure  5).  In  contrast,  the  behavior  required  by  the 
already  increases  by  ∆fin  ≈  19  Pg/yr  or  2.4  ppm/yr  for  an
Conservation  Law  decays  as  fast  or  faster  than  that  of  the 
ave age  temperature  incline  of  1°C.  An  even  stronger 
fastest sink (see (28)). 
variation  can  be  expected  for  the  land  vegetation  with  an 
The  observed  decay  of  14C  shows  that  the  corresponding 
increased decomposition and reduced uptake of CO2 at rising 
absorption is determined by a single decay time and operates 
temperature (Lee [34]; Salby [11]).  
on a time scale of only about one decade (see Figure 5). This 
Together this causes an incline of the atmospheric CO2 level 
scale  is  the  same  for  the  natural  carbon  cycle  as  for  the 
which  is  larger  than  all  apparent  human  activities,  but  its 
anthropogenic cycle. Therefore, it is unrealistic to differentiate 
contribution is completely neglected in the official accounting 
between a residence time and different adjustment times   
schemes.  
In this context it should be noticed that due to re-emissions 
Also  melting  permafrost  and  emissions  of  volcanoes  on 
of 14CO2 from extraneous reservoirs the real residence time of
land and under water as well as any emissions at earthquakes 
14CO2  in  the  atmosphere  as  well  as  that  of  the  other 
are  not  considered.  In  addition,  actual  estimates  of  dark 
isotopologues of CO2 can only be shorter, ev n shorter than a 
respiration  suggest  that  under  global  warming  conditions 
decade (for details see subsection 5.7.3 and App ndix B). 
INFORMATION 
whole-plant  respiration  could  be  around 30%  higher  than 
existing estimates (Huntingford et al. [35]). This longer list of 
5.6. Temperature Dependence 
different native events and effects is completely embezzled in 
RELEASED UNDER THE 
According  to  (9)  or  (10)  we  see  that  with  increasing 
the favored IPCC models.  
atmospheric concentration over the Industrial Era from 280 to 
Equally  inconsistent  is  the  presumption  that  additional 
400  ppm  either  the  residence  time  must  be  increased  with 
uptake of anthropogenic CO2, which represents less than 1% 
temperature  from  3  to  about  4  yr,  or  τ
of  the  total  over  the  Industrial  Era,  has,  somehow,  exceeded 
R  is  considered  to  be
constant and the total emissions were rising from 93 to about 
the  storage  capacity  of  oceans  and  other  surface  and 
130 ppm/yr, synchronously increasing the concentration. Both 
sub-surface reservoirs, capacity which is orders of magnitude 
these  limiting  cases  are  in  agreement  with  a  temperature 
greater.  A  reduced  absorption  is  rather  the  consequence  of 
anomaly of about 1.2 °C over this period (see GISS [9]), when 
global warming than of saturation. Due to Henry's law and its 
 OFFICIAL 
we assume the  maximum temperature coefficients βτ = 0.74
temperature  dependence  not  only  the  partial  pressure  in  sea 
yr/°C  or  β
water  increases,  but  also  the  solubility  of  CO
e  =  24  ppm/yr/°C.  However,  generally  both 
2  in  water 
temperature induced natural emissions as well as temperature 
declines exponentially with temperature and, thus, reduces the 
dependent  absorptions  together  will  dictate  the  inclining 
CO2 uptake. Often is this effect incorrectly misinterpreted as 
concentration in the atmosphere. 
saturation caused by a limited buffer capacity and dependent 
In  any  way,  as  we  see  from  Figure  8,  is  the  CO
on  the  concentration  level.  But  here  we  consider  an  uptake 

concentration  dominantly  empowered  by  the  temperature 
changing  with  temperature,  as  this  is  known  for  chemical 
increase;  with  only  one  unique  decay  process  not  human 
reactions,  where  the  balance  is  controlled  by  temperature. 
activities but almost only natural impacts have to be identified 
How  strongly  the  biological  pump  (see  Appendix  A)  and 


ACT 1982
INFORMATION 
RELEASED UNDER THE 
 OFFICIAL 

Earth Sciences 2019; 8(3): 139-159 
152 
item). 
half of the emissions remained in the atmosphere since 1750
Since  the  fossil  fuel  emissions  have  a  leaner  difference 
and  "the  removal  of  all  the  human-emitted  CO2  from  the 
(δ13C)fuel-atm  =  -18  ‰  compared  to  the  atmosphere,  or
atmosphere  by  natural  processes  will  take  a  few  hundred 
(δ13C)fuel-VPDB = -25 ‰ with respect to the international VPDB
thousand  years  (high  confidence)"  (see  AR5  [1],  Chap. 
carbonate standard (Coplen [38]), the rising human emissions 
6-Summary and Box 6.1) can be simply refuted by the isotope
over the 30 yr interval can only have contributed to a decline 
measurements  at  Mauna  Loa.  If  the  113  ppm  CO2  increase
of  ∆  =  (δ13C)
×
fuel-atm 1.8%  =  -18‰×1.8%  =  -0.32  ‰  or  a
since 1750 (28.8% of the present concentration of 393 ppm -
(δ13C)atm  =  -7.92‰  in  2010.  Thus,  the  difference  to  -8.3‰,
average  between  2007  and  2016)  would  only  result  from
which  is  more  than  50%,  in  any  case  must  be  explained  by 
human impacts and would have cumulated in the atmosphere,
other effects.  
the  actual  (δ13C)atm  value  should  have  dropped  by  ∆  =
One possible explanation for a faster decline of (δ13C)
×
atm to
(δ13C)fuel-atm 28.8%  =  -18‰×28.8%  =  -5.2‰  to  (δ13C)atm  ≈
-8.3‰ can be - even with oceans as source and an 13C/12C ratio
-7‰ -5.2‰ = -12.2‰, which by far is not observed. (δ13C)atm
in  sea  water  greater  than  in  air  (particularly  in  the  surface
in 1750 was assumed to have been -7‰.
layer) - that the lighter 12CO2 molecules are easier emitted at
the ocean's surface than 13CO
5.7.3. Fossil Fuels are Devoid of Radiocarbon 
2, this with the result of a leaner
13C concentration in air and higher concentration in the upper 
“Because  fossil  fuel  CO2  is  devoid  of  radiocarbon  (14C),
water  layer  (see  also:  Siegenthaler  &  Münnich  [39]).  From 
reconstructions  of  the  14C/C  isotopic  ratio  of  atmospheric 
water we also know that its isotopologues are evaporated with 
CO2  from  tree  rings  show  a  declining  trend,  as  expected 
slightly different rates.  
from the addition of fossil CO2 (Stuiver and Quary, 1981; 
Such  behavior  is  in  agreement  with  the  observation  that 
Levin et al., 2010)  Yet nuclear weapon tests  in the 1950s 
with  higher  temperatures  the  total  CO
and 1960s have been offsetting that declining trend signal 
2  concentration  in  the 
atmosphere  increases,  but  the  relative  13CO
by adding 14C to the atmosphere. Since this nuclear weapon 
2  concentration
ACT 1982
decreases. This can be observed, e.g., at El Niño events (see: 
induced  14C  pulse  in  the  atmosphere  has  been  fading,  the 
1

M. L. Salby [40], Figure 1.14; Etheridge et al. [41]; Friedli et
C/C  isotopic  ratio  of  atmospheric  CO2  is  observed  to 
al. [42]).
resume  its  declining  trend  (Naegler  and  Levin,  2009; 
We also remind at the Mauna Loa curve, which shows for 
Graven et al., 2012).” 
the total emissions a seasonal variation with an increasing CO2 
For 14C we can adduce almost the same comments as listed 
concentration from about October till May and a decline from 
for 13C. Fossil CO2 devoid of 14C will reduce the 14C/C ratio of
June to September. The increase is driven by respiration and 
the  atmosphere,  this  is  valid  for  our  approach  in  the  same 
decomposition  mainly  on  the  Northern  Hemisphere  (NH)  as 
manner  as  for  the  IPCC  schemes.  But,  as  no  specific 
well as the temperature on the Southern Hemisphere (SH) and 
accumulation  of  anthropogenic  molecules  is  possible 
also  local  temperature  effects.  The  (δ13C)atm  value  is  just
(equivalence  principle),  this  decline  can  only  be  expected 
anti-cyclic  to  the  total  CO2  concentration  (AR5  [1],  Figure 
proportional  to  the  fraction  of  fossil  fuel  emission  to  total 
6.3) with a minimum at maximum CO2 concentration and with 
emission. Before 1960 this was not more than 1% and actually 
seasonal variations of 0.3 - 0.4‰, the same order of magnitude 
it is about 4.3%. 
as the fossil fuel effect. 
14C is continuously  formed in the upper atmosphere from 
An increase of 13C in the upper strata of oceans also results 
14N  through  bombardment  with  cosmic  neutrons,  and  then 
INFORMATION 
from  an  increased  efficiency  of  photosynthesis  for  lighter 
rapidly  oxidizes  to  14CO2.  In  this  form  it  is  found  in  the
CO2.  Plankton  accumulates  this  form  and  sinks  to  lower 
atmosphere  and  enters  plants  and  animals  through 
layers, where it decomposes and after longer times is emitted 
photosynthesis and the food chain. The isotopic 14C/C ratio in 
RELEASED UNDER THE 
in  higher  concentrations  with  stronger  upwelling  waters 
air  is  about  1.2⋅10-12,  and  can  be  derived  either  from  the
particularly in the Eastern Tropic Pacific. It is also known that 
radioactivity  of  14C,  which  with  an  average  half-lifetime  of 
the 13C concentrations are by far not equally distributed over 
5730 yr decays back to 14N by simultaneously emitting a beta 
the Earth's surface. Thus  it can be expected that with volcanic 
particle,  or  by  directly  measuring  the  amount  of  14C  in  a 
and tectonic activities different ratios will be released. 
sample by means of an accelerator mass spectrometer. 
So,  without  any  doubts  fossil  fuel  emissions  will  slightly 
Fossil fuels older than several half-lives of radiocarbon are, 
dilute the 13CO2 conc ntration in air. But presupposing regular
thus, devoid of the 14C isotope. This influence on radiocarbon 
conditions for the uptake process (equivalence principle) they 
measurements is  known  since the investigations of H. Suess 
 OFFICIAL 
contribute  less  than  50%  to  the  observed  decrease.  The 
[43] who observed a larger 14C decrease (about 3.5%) for trees
difference  has  to  be  explained  by  additional  biogeochemical 
from  industrial  areas  and  a  smaller  decline  for  trees  from
processes. Particularly the seasonal cycles and events like El 
unaffected  areas.  This  so-called  Suess  or  Industrial  effect  is
Niños  are  clear  indications  for  a  stronger  temperature 
important  for  reliable  age  assignments  by  the  radiocarbon
controlled modulation of the  (δ13C)atm value. Therefore is an
method  and  is  necessary  for  respective  corrections.  But  for
observed decline of the 13C/12C ratio over recent years by far 
global climate considerations  it gives no new  information, it
not  a  confirmation  of  an  anthropogenic  global  warming 
only  confirms  the  calculations  based  on  the  human  to  total
(AGW) theory.  
emission  rate  (see  above),  and  it  clearly  shows  that  an
Also  the  widely  spread  but  wrong  declaration  that  "about 
assumed  accumulation  of  anthropogenic  CO2  in  the

153 
Hermann Harde: What Humans Contribute to Atmospheric CO2: Comparison of Carbon Cycle Models with Observations 
atmosphere contradicts observations. 
completely sequestered beneath the Earth's surface by a single 
More  important  for  climate  investigations  is  that  after  the 
absorption process. A substantial fraction is therefore returned 
stop  of  the  nuclear  bomb  tests  1963  14C  could  be  used  as  a 
to  the  atmosphere  through  re-emission  (e.g.,  through 
sensitive  tracer  in  the  biosphere  and  atmosphere  to  study 
decomposition of vegetation which has absorbed that 14C), and 
temporal carbon mixing and exchange processes in the carbon 
in  average  it  takes  several  absorption  cycles  to  completely 
cycle. As the bomb tests produced a huge amount of thermal 
remove  that  14C  from  the  atmosphere.  This  simply  modifies 
neutrons  and  almost  doubled  the  14C  activity  in  the 
the effective absorption for radiocarbon, but with a resulting 
atmosphere, with the end of these tests the temporal decline of 
decay  which remains exponential (see Figure 5). Unlike any 
the excess radiocarbon activity in the atmosphere can well be 
dilution  effect  by  fossil  fuel  emission,  which  is  minor  (see 
studied. This decline is almost completely independent of the 
Appendix B), this re-emission slows decay over what it would 
radioactive  lifetime,  but  practically  only  determined  by  the 
be in the presence of pure absorption alone. Therefore is the 
uptake through extraneous reservoirs.  
apparent absorption time - as derived from the 14C decay curve 
Such  decline  has  already  been  displayed  in  Figure  5  as 
- longer than the actual absorption time.
fractionation-corrected ‰-deviations ∆14CO2 from the Oxalic
In  this  context  we  emphasize  that  apart  from  some  minor
Acid  activity  corrected  for  decay,  this  for  a  combination  of 
influence  due  to  fractionation  all  CO2  isotopologues  are 
measurements  at  Vermunt  and  Schauinsland  (Magenta  Dots 
involved in the same multiple re-emission cycles. But in (23) 
and Green Triangles; data from Levin et al. [17]). The decay is 
or (32) this is already cons dered in the total balance via the 
well represented by a single exponential with a decay constant 
emission rates, for which it makes no difference, if the same or 
of  about  15  yr  (Dashed  Blue).  For  similar  observations  see 
meanwhile  exchanged  molecules 
re  recycled  to  the 
also Hua et al. [18] and Turnbull et al. [19]. Thus, the decay 
atmosphere.  In  contrast  to  this  are  14CO2  isotopologues 
satisfies the relation  
identified  through  their  radioactivity,  and  in  the  worst  case 
without  any  dilution  or  exchange  processes  in  an  external 
ACT 1982
dC'
1
14 = −
C' , 
 (31) 
reservoir τ
14
14  would  approach  the  radioactive  lifetime.  On  the
dt
τ14
other h nd, at strong diffusion, dilution or sequestration of 14C 
in such reservoirs  τ14 would converge to τR. Consequently it
where C'14 represents the excess concentration of radiocarbon 
fol ws from the observed 14C decay shown in Figure 5 that 
above  a  background  concentration  in  the  atmosphere.  It 
this provides an upper bound on the actual absorption time τR,
corresponds to absorption that is proportional to instantaneous 
which can be only shorter. Both are tremendously shorter than 
concentration  with  an  apparent  absorption  time  τ14  slightly
the adjustment time requested by the IPCC.  
more than a decade. 
The  exponential  decay  of  14C  with  only  one  single  decay 
Because CO2 is conserved in the atmosphere  it can change 
time  proves  models  with  multiple  relaxation  times  to  be 
only through an imbalance of the surface fluxes eT and aT. This 
wrong.  At  the  same  time  it  gives  strong  evidence  for  a  first 
holds for all isotopologues of CO2 in the same way. For this 
order absorption process as considered in Section 4.2  
reason,  its  adjustment  to  equilibrium  must  proceed  through 
those influences. They are the same influences that determine 
5.7.4. Higher Fossil Fuel Emissions in the Northern 
the  removal  time  of  CO2  in  the  atmosphere.  If  CO2  is 
Hemisphere 
perturbed  impulsively  (e.g.,  through  a  transient  spike  in 
“Most  of  the  fossil  fuel  CO2  emissions  take  place  in  the 
emission),  its  subsequent  decay  must  track  the  removal  of 
industrialised  countries  north  of  the  equator.  Consistent 
INFORMATION 
perturbation  CO2,  C',  which  in  turn  is  proportional  to  its 
with  this,  on  annual  average,  atmospheric  CO2 
instantaneous  concentration.  Determined  by  the  resulting 
measurement stations in the NH record increasingly higher 
RELEASED UNDER THE 
imbalance  between  eT  and  aT,  that  decay  is  governed  by  the 
CO2 concentrations than stations in the SH, as witnessed by 
perturbation form of the balance equation: 
the observations from Mauna Loa, Hawaii, and the South 
dC'
1
Pole (see Figure 6.3). The annually averaged concentration 
= −
C' , 
 (32) 
difference  between  the  two  stations  has  increased  in 
dt
τR
proportion  of  the  estimated  increasing  difference  in  fossil 
which is the same form as the observed decay of 14C following 
fuel combustion emissions between the hemispheres (Figure 
elimination of the pe turbing nuclear source. But there is still 
6.13;  Keeling  et  al.,  1989;  Tans  et  al.,  1989;  Fan  et  al., 
one important difference between these equations.  
1999)”. 
 OFFICIAL 
Eq.(32) is the perturbation form of (23) with a decay time 
The  strongest  terrestrial  emissions  result  from  tropical 
τR, the residence time, because 1/τR describes the rate at which
forests, not industrial areas. The strongest oceanic emissions 
CO2 is removed from the atmosphere, this as the result of the 
can be seen from the  map of Takahashi et al. [32]. They are 
balance between all absorption and emission processes.  
In  contrast  to  this  describes  (31)  a  decay  process,  which 
implicitly also considers some back-pumping of radiocarbon 
2 A calculation similar to Figure 8 but with a residence time of 15 yr as an upper 
to the atmosphere (see Appendix B, (37)). So, from all 14C that 
bound would require to reduce the natural emissions at pre-industrial times from 93 
is removed from the atmosphere with the time constant τ
ppm/yr to 19 ppm/yr. Then the anthropogenic contribution would supply 59 ppm
R - in
which is 15% of the total atmospheric concentration or 52% of the increase since 
the same way as all isotopes -, only some smaller fraction is 
1850. 

Earth Sciences 2019; 8(3): 139-159 
154 
between  10°N  and  10°S  in  the  Eastern  Tropic  Pacific. 
by  a  single  balance  equation,  the  Conservation  Law  (23), 
Nevertheless,  there  is  no  doubt  that  industrial  emissions 
which considers the total atmospheric CO2 cycle, consisting of 
endow  their  fingerprints  in  the  atmosphere  and  biosphere 
temperature  and  thus  time  dependent  natural  emissions,  the 
(Suess effect). The influence and size of these emissions has 
human activities and a temperature dependent uptake process, 
already  been  discussed  above,  and  their  different  impact  on 
which scales proportional with the actual concentration. This 
the  two  hemispheres  can  be  estimated  from  Figure  6.3c  of 
uptake  is  characterized  by  a  single  time  scale,  the  residence 
AR5 [1] indicating a slightly faster decline of (δ13C)atm for the
time  of  about  3  yr,  which  over  the  Industrial  Era  slightly 
NH  in  agreement  with  predominantly  located  industrial 
increases  with temperature. Only this concept is in complete 
emissions  in  this  hemisphere.  Even  more  distinctly  this  is 
conformity  with  all  observations  and  natural  causalities.  It 
illustrated by Figure 6.13 of AR5 [1] for the difference in the 
confirms  previous  investigations  (Salby  [7,  10];  Harde  [6]) 
emission  rates  between  the  northern  and  SH  with  8 PgC/yr, 
and shows the key deficits of some widespread but largely ad 
which can be observed as a concentration difference between 
hoc  carbon  cycle  models  used  to  describe  atmospheric  CO2, 
the  hemispheres  of  3.8  ppm.  But  this  is  absolutely  in  no 
failures which are responsible for the fatal conclusion  hat the 
dissent to our result in Section 4 that from globally 4.7 ppm/yr 
increase  in  atmospheric  CO2  over  the  past  270  years  is 
FFE and LUC (average emission over 10 yr) 17 ppm or 4.3 % 
principally anthropogenic. 
contribute  to  the  actual  CO2  concentration  of  393  ppm 
For a conservative assessment  we  find from Figure 8 that 
(average).  This  impact  is  of  the  same  size  as  seasonal 
the anthropogenic contribu ion to the observed CO2 increase 
variations  observed  at  Mauna  Loa  before  flattening  and 
over  the  Industrial  Era  is  significantly  less  than  the  natural 
averaging the measurements. 
influence.  At  equilibrium  this  contribution  is  given  by  the 
fraction of human to native impacts. As an average over the 
5.7.5. Human Caused Emissions Grew Exponentially 
period  2007-2016  the  anthropogenic  emissions  (FFE&LUC 
“The rate of CO2 emissions from fossil fuel burning and land 
together)  d nated  not  more  than  4.3%  to  the  total 
ACT 1982
use  change  was  almost  exponential,  and  the  rate  of  CO2 
concentration  of  393  ppm,  and  their  fraction  to  the 
increase in  the atmosphere  was also almost  exponential and 
atmospheric increase since 1750 of 113 ppm is not more than 
about half that of the emissions, consistent with a large body of 
17  ppm  or  15%.  With  other  evaluations  of  absorption,  the 
evidence about changes of carbon inventory in each reservoir 
con ribution  from  anthropogenic  emission  is  even  smaller. 
of the carbon cycle presented in this chapter”. 
Thus,  not  really  anthropogenic  emissions  but  mainly  natural 
The size and influence of FFE and LUC on the atmospheric 
processes, in particular the temperature, have to be considered 
CO
as the dominating impacts for the observed CO
2  concentration  has  extensively  been  discussed  in  the 
2 increase over 
preceding sections. Only when violating fundamental physical 
the last 270 yr and also over paleoclimate periods.  
principles  like  the  equivalence  principle  or  denying  basic 
causalities  like  a  first  order  absorption  process  with  only  a 
Acknowledgements 
single  absorption  time,  the  CO2  increase  can  be  reproduced 
with anthropogenic emissions alone. 
The author thanks Prof. Murry Salby, formerly Macquarie 
In contrast to that we could demonstrate that conform with 
University  Sydney,  for  many  stimulating  discussions  when 
the  rising  temperature  over  the  Industrial  Era  and  in 
preparing  the  paper,  and  Jordi  López  Fernández,  Institute  of 
conformity  with  all  physical  legalities  the  overwhelming 
Environmental Assessment and Water Studies Barcelona, for 
INFORMATION 
fraction of the observed CO
his support when searching for temperature data.  
2 increase  has to be explained by 
native  impacts.  Such  simulations  reproduce  almost  every 
This  research  did  not  receive  any  specific  grant  from 
detail of the observed atmospheric CO
funding agencies in the public, commercial, or not-for-profit 
RELEASED UNDER THE 
2 increase (see Figures 8 
and 10). And from observations of natural emissions it can be 
sectors. 
seen  that  they  are  increasing  slightly  exponential  with 
temperature (Takahashi et al. [32]; Lee [34]).  
Appendix 
Thus, no one of the preceding lines of evidence can really 
support the above statement that "fossil fuel burning and land 
Appendix A 
use change are the dominant cause of the observed increase in 
The absorption efficiency of extraneous reservoirs has been 
atmospheric  CO2  concentration."  In  fact,  they  apply  in  the 
claimed  to  have  decreased,  based  on  changes  in  the 
same  way  for  our  concept,  and  thus  they  are  useless  to 
arbitrarily-defined airborne fraction (e.g., Le Quéré et al. [12]; 
 OFFICIAL 
disfavour  our  approach.  The  isotopic  studies  rather  confirm 
Canadell  et  al.  [44]).  Such  claims  are  dubious  because  they 
our  ansatz  of  a  first  order  absorption  process  with  a  single 
rely on the presumption that changes of CO
absorption  time,  which  is  significantly  shorter  than  one 
2 are exclusively of 
anthropogenic origin. Nor are the claims supported by recent 
decade, and they refute the idea of cumulating anthropogenic 
atmospheric  CO
emissions in the atmosphere.  
2  data.  Gloor  et  al.  [45]  found  that  decadal 
changes  of  AF  followed  from  changes  in  the  growth  of 
anthropogenic  emissions  -  not  from  changes  in  absorption 
6. Conclusion
efficiency,  which  were  comparatively  small.  Further, 
uncertainties  in  emission  and  absorption  exceeded  any 
The increase of CO2 over recent years can well be explained

155 
Hermann Harde: What Humans Contribute to Atmospheric CO2: Comparison of Carbon Cycle Models with Observations 
changes  in  AF.  Ballantyne  et  al.  [46]  arrived  at  a  similar 
atmosphere.  Anthropogenic  CO2  in  surface  water  is  then 
conclusion. They used global atmospheric CO2 measurements 
quickly  removed.  It  is  also  well  known  that  higher  concen-
and  CO2  emission  inventories  to  evaluate  changes  in  global 
trations of CO2 magnify photosynthesis. At increased atmos-
CO2  sources  and  sinks  during  the  past  50  years.  Their  mass 
pheric  CO2,  the  plankton  community  consumed  39%  more 
balance  analysis  indicates  that  net  CO2  uptake  significantly 
DIC (Riebesell et al. [53]). During summer and autumn, sur-
increased,  by  about  0.18  Pg/yr  (0.05  GtC/yr)  and,  between 
face CO2 can rapidly increase to 1000 ppm - more than twice 
1960 and 2010, that global uptake actually doubled, from 8.8 
the  concentration  of  CO2  in  the  atmosphere.  Surface  water 
to 18.4 Pg/yr. It follows that, without quantitative knowledge 
then  significantly  enhances  natural  emission  to  the  atmos-
of  changes  in  natural  emission,  interpretations  based  on  AF 
phere.  Conversely,  during  winter,  surface  CO2  remains  at 
are little more than speculative. 
about  340  ppm.  Despite  reduced  photosynthesis,  CO2  in 
The uptake and outgassing of atmospheric CO2 by oceans is 
surface  water  then  remains  below  equilibrium  with  the 
simulated  with  complex  marine  models.  How  much  CO2 
atmosphere,  reflecting  efficient  removal  through  downward 
enters  or  leaves  the  ocean  surface  is  calculated  from  the 
transport by the biological pump. It  is  noteworthy that these 
difference between atmospheric and surface concentrations of 
strong seasonal variations of CO2 in surface water are mani-
CO2, modified by the Revelle factor. However, most of these 
fest in the record of atmospheric CO2 (see Figures 9 and 10). 
models involve assumptions which are not in agreement with 
Under  steady  state  conditions,  diffusion  of  CO2  into  the 
observed  behavior  (see,  e.g.,  Steele  [47]).  They  assume  that 
ocean is believed to require about 1 year to equilibrate with an 
the  surface  layer  absorbs  CO2  through  equilibrium  with 
atmospheric  perturbation.  But,  when  increased  sunlight 
atmospheric  concentration.  On  this  premise,  they  calculate 
enhances  photosynthesis,  such  equilibration  is  no  longer 
how much Dissolved Inorganic Carbon (DIC) will be added to 
achieved.  Perturbation  CO2  is  then  simply  transported  to 
the ocean based on increased atmospheric CO2 since pre-indu- 
depth,  where  it  is  sequestered  from  surface  waters 
strial times. In reality, the surface layer is not at equilibrium 
(McDonnell et al. [54]). Under such conditions uptake of CO2 
ACT 1982
with  the  atmosphere.  A  difference  in  concentration  results 
is  not  restricted  by  the  Revelle  factor  but  by  the  biological 
from  conversion  of  CO2  into  organic  carbon  by 
pump.  
photosynthesis. Organic carbon produced then sinks into the 
The  foregoing  processes  are  controlled  essentially  by 
deep ocean, where it is sequestered. This downward transport 
sunlight and temperature. There is no reason to believe that net 
to  the  deep  ocean  is  known  as  the  biological  pump.  In  the 
primary production, the biological pump, and sequestration of 
Northeastern Atlantic basin, e.g., Benson et al. [33] report on 
CO2  below  surface  waters  would  be  the  same  today  as  270 
seasonal  pressure  differences  between  the  ocean  and 
years ago, when temperature and atmospheric CO2 were likely 
atmosphere of ∆pCO2 = -70 µatm and an air-sea CO2 flux of
lower. 
220 g/m2/yr. Only in those regions where strong upwelling of 
In  simulating  transport  of  carbon  in  the  ocean,  complex 
DIC from the deep ocean exceeds sequestration of carbon via 
models assume behavior that  is  found in tracers like chloro-
photosynthesis can CO2 be outgassed to the atmosphere. The 
fluorocarbons (CFCs). Because those species accumulate near 
latter is found primarily in the tropical oceans (Takahashi et al. 
the ocean surface, models assume DIC does as well. But un-
[32]; Zhang et al. [31]). Several models es imate that, without 
like  CFCs,  which  are  inert,  CO2  entering  sunlit  waters  is 
the  biological  pump,  atmospheric  CO2  would  be  200  to  300 
quickly converted to organic matter by photosynthesis (Steele 
ppm higher than current levels (see also Evans [48]).  
[47]).  Although  dissolved  CFCs  and  dissolved  carbon  are 
INFORMATION 
With increasing primary production, carbon export to depth 
passively transported in the same manner, particulate organic 
also grows. Arrigo et al. [49] reported that  since 1998, annual 
carbon  (alive  or  dead)  behaves  very  differently.  It  rapidly 
primary  production  in  he  Arctic  has  increased  by  30%. 
sinks,  removing  carbon  from  surface  water  through  mecha-
RELEASED UNDER THE 
Steinberg et al. [50] observed a 61% increase in meso-plank-
nisms which do not operate on CFCs.  
ton  between  1994  and  2006  in  the  Sargasso  Sea.  The  North 
The removal of carbon from surface water depends on the 
Atlantic  coccolithophores  h ve  increased  by  37%  between 
sinking  velocity  and  also  on  how  rapidly  organic  matter  is 
1990 and 2012 (Krumhardt et al. [51]). And Chavez et al. [52] 
decomposed. After descending below the pycnocline (depths 
found  a  dramatic  increase  in  primary  production  in  the  Peru 
of  500-1000  meters),  carbon  is  effectively  sequestered  - 
Current since the end of the Little Ice Age (LIA). Together, the 
because water at those depths does not return to the surface for 
increase  in  primary  production  and  downward  transport  of 
centuries  (Weber  et  al.  [55]).  For  the  atmosphere,  this 
organic carbon is sufficient to account for anthropogenic CO2 
long-term  sequestration  translates  into  removal  that  is 
 OFFICIAL 
that was absorbed from the atmosphere (Steele [47]). 
effectively  permanent.  Before  such  carbon  can  return  to  the 
Further, seasonal changes in surface CO2 illustrate that ab-
atmosphere,  fossil  fuel  reserves  will  have  long  since  been 
sorption of CO2 by the oceans and accumulation of DIC near 
exhausted.  
the surface are determined, not by the Revelle factor, but by 
The  combination  of  sinking  velocities  and  sequestration 
the biological pump. Evans et al. [48] found from buoy data 
depth  suggests  that  a  significant  fraction  of  primary  produc-
off the coast of Newport, Oregon that each spring photosyn-
tion is sequestered in a matter of days to weeks (Steele [47]). 
thesis  lowers  ocean  surface  CO2  to  200  ppm  -  far  below 
Therefore,  increasing  primary  production  leads  to  a  propor-
current atmospheric concentrations and much lower than what 
tionate increase and rapid export of carbon to depth. If marine 
would  be  expected  from  equilibrium  with  a  pre-industrial 
productivity  has  increased  since  pre-industrial  times,  it  will 


ACT 1982
INFORMATION 
RELEASED UNDER THE 
 OFFICIAL 

157 
Hermann Harde: What Humans Contribute to Atmospheric CO2: Comparison of Carbon Cycle Models with Observations 
Primed  quantities  are  now  referenced  against  unperturbed 
[10] M. L. Salby, "Relationship Between Greenhouse Gases and
values  before  introduction  of  the  nuclear  source.  From  a 
Global Temperature", Video Presentation, April 18, 2013.
Helmut-Schmidt-University Hamburg
balance  for  the  Earth  layer  it  follows  that  in  good 
https://www.youtube.com/watch?v=2ROw_cDKwc0.
approximation  e'14  opposes  the  atmospheric  absorption  rate 
C'
[11] M.  L.  Salby,  "What  is  Really  Behind  the  Increase  of
14/τR  minus  the  sequestration  rate  C'E,14/τ14,  for  which  it  is
assumed  that  the  concentration  in  the  upper  layer  C'
Atmospheric CO2"? Helmut-Schmidt-University Hamburg, 10.
E,14  is 
October 2018, https://youtu.be/rohF6K2avtY
almost the same as the concentration C'14 in the atmosphere. 
Thus,  re-emission  simply  modifies  the  effective  absorption, 
[12] C. Le Quéré, M. R. Raupach, J. G. Canadell, G. Marland et al.,
which for 14C is controlled by the apparent absorption time τ
"Trends  in  the  sources  and  sinks  of  carbon  dioxide",  Nature
14
and not the residence time τ
Geosci., 2, pp. 831–836, 2009. doi:10.1038/ngeo689.
R in agreement with (34).
Unlike the dilution effect, which is minor, this slows decay 
[13] P. Tans, NOAA/ESRL and R. Keeling, Scripps Institution of
over what it would be in the presence of absorption alone. The 
Oceanography (scrippsco2.ucsd.edu/), 2017.
apparent  absorption  time  is  therefore  longer  than  the  actual 
https://www.esrl noaa.gov/gmd/ccgg/trends/data html.
absorption  time,  which  must  even  be  shorter  than  a  decade. 
[14] F. Joos, M. Bruno, R. Fink, U. Siegenthaler, T. F  Stocker, C. Le
Integration  of  (37)  or  (34)  exactly  reproduces  a  pure  expo-
Quéré, J. L. Sarmiento, "An efficient and accurate representa-
nential decay in Figure 13 with an e-folding time τ14 =15 yr.
tion  of  complex  oceanic  and  biospheric  models  of  anthropo-
genic  carbon  uptake",  Tellus  B  48,  pp.  397–417,  1996.
doi:10.1034/j.1600-0889.1996.t01 2-00006 x.
References 
[15] J.  Hansen,  M.  Sato,  P.  Kharecha,  G.  Russell,  D.  W.  Lea,  M.
Siddall, "Climate change and trace gases", Phil. Trans. R. Soc.
[1]
AR5,  In:  Stocker,  T.  F.,  Qin,  D.,  Plattner,  G.-K.,  Tignor,  M.,
A 365, pp. 1925–1954, 2007. doi:10.1098/rsta.2007.2052.
Allen,  S.  K.,  Boschung,  J.,  Nauels,  A.,  Xia,  Y.,  Bex,  V.,
Midgley,  P.  M.  (Eds.),  "Climate  Change  2013:  The  Physical
[16] J. Hansen  M. Sato  G. Russell, K. Pushker, "Climate
ACT 1982
Science  Basis.  Contribution  of  Working  Group  I  to  the  Fifth
sensitivity, sea level,  nd atmospheric CO2", Philos. Trans. R.
Assessment Report of the Intergovernmental Panel on Climate
Soc. A, 371, 20120294, 2013. doi:10.1098/rsta.2012.0294.
Change",  Cambridge  University  Press,  Cambridge,  United
https://www.nasa.gov/
Kingdom and New York, NY, USA, 2013.
[17] I.  Levin,  B   Kromer,  and  S.  Hammer,  "Atmospheric  ∆14CO2
[2]
C. Le Quéré et al., "Global Carbon Budget 2017", Earth Syst.
trend in Western European background air from 2000 to 2012",
Sci.  Data  Discuss.,  https://doi.org/10.5194/essd-2017-123
Tellus B 65, pp. 1-7, 2013.
Open Access Earth System Science Data Discussions, Manu-
script under review for journal Earth Syst. Sci. Data, 2017
[18] Q. Hua, M. Barbetti, A. Z. Rakowski, "Atmospheric radiocar-
bon  for  the  period  1950–2010".  RADIOCARBON  55,  pp.
[3]
CICERO, Center for International Climate Research, Oslo, R
2059–2072,  (2013).  Supplementary  Material  Table  S2c,
Andrew: http://folk.uio no/roberan/GCP2017.shtml  2017.
https://doi.org/10.2458/azu_js_rc.v55i2.16177
[4]
CDIAC,  2017:  Carbon  Dioxide  Information  Analysis  Center
[19] J. C. Turnbull, S. E. Mikaloff Fletcher, I. Ansell, G. W. Brails-
http://cdiac.ornl.gov/trends/emis/glo_2014.html.
ford, R. C. Moss, M. W. Norris, K. Steinkamp, "Sixty years of
radiocarbon dioxide measurements at Wellington, New
[5]
C. D. Keeling, S. C. Piper, R. B. Bacastow, M. W hlen, T. P.
Zealand: 1954–2014", Atmos. Chem. Phys. 17, pp. 14771–
Whorf,  M.  Heimann,  H.  A.  Meijer,  "Atmospheric  CO2  and
14784, 2017. https://doi.org/10.5194/acp-17-14771-2017.
13CO2 exchange with the terrestrial biosphere and oceans from
INFORMATION 
1978 to 2000: Observation  and carbon cycle implications", In:
[20] U. Siegenthaler, J. L. Sarmiento, "Atmospheric carbon dioxide
Ehleringer,  J.  R.,  Cerling,  T.  E.,  Dearing   M.  D.  (Eds.),  A 
and the ocean", Nature 365, pp. 119-125, 1993.
History  of  Atmospheric  CO2  and  Its  Effects  on  Plants,
RELEASED UNDER THE 
Animals, and Ecosystems. Springer Science+Business Media,
[21] P. Dietze, IPCC's Most Essential Model Errors, 2001.
New  York,  NY,  USA,  and  Heidelberg,  Germany,  pp.  83–113
http://www.john-daly.com/forcing/moderr.htm; (Carbon Model
(actualized by Scripps-Institutes, USA), 2005.
Calculations, http://www.john-daly.com/dietze/cmodcalc htm).
[6]
H. Harde, "Scrutinizing the carbon cycle and CO
[22] G.  C.  Cawley,  "On  the  Atmospheric  Residence  Time  of
2 residence
time in the atmosphere"  Global and Planetary Change 152, pp.
Anthropogenically Sourced Carbon Dioxide", Energy Fuels 25,
19–26, 2017.
pp. 5503–5513, 2011. dx.doi.org/10.1021/ef200914u
http://dx.doi.org/10.1016/j.gloplacha.2017.02.009.
[23] H.-J. Lüdecke, C. O. Weiss, "Simple Model for the
[7]
M. L. Salby, "Atmospheric Carbon", Video Presentation, July
Anthropogenically Forced CO2 Cycle Tested on Measured
 OFFICIAL 
18, 2016. University College London.
Quantities", JGEESI, 8(4), pp. 1-12, 2016. DOI:
https://youtu.be/3q-M_uYkpT0.
10.9734/JGEESI/2016/30532.
[8]
P. Köhler, J. Hauck, C. Völker, D. A. Wolf-Gladrow, M. Butzin,
[24] R. E. Essenhigh, "Potential dependence of global warming on
J. B. Halpern, K. Rice, R. E. Zeebe, Comment on “Scrutinizing
the residence time (RT) in the atmosphere of anthropogenically
the carbon cycle and COresidence time in the atmosphere” by
sourced  carbon  dioxide",  Energy  Fuel  23,  pp.  2773–2784,
H. Harde, Global and Planetary Change 164, pp. 67-71, 2017.
2009. http://pubs.acs.org/doi/abs/10.1021/ef800581r.
https://doi.org/10.1016/j.gloplacha.2017.09.015
[25] E. Berry, "Human CO2 has little effect on atmospheric CO2", 2019.
[9]
GISS, 2017: Goddard Institute for Space Studies:
https://edberry.com/blog/climate-physics/agw-hypothesis/contradi
https://data.giss nasa.gov/gistemp/.
ctions-to-ipccs-climate-change-theory/

Earth Sciences 2019; 8(3): 139-159 
158 
[26] NOAA, 2017:
[39] U. Siegenthaler, K. O. Münnich, "13C/12C fractionation during
https://www.esrl noaa.gov/psd/data/gridded/data ncep reanalys
CO2 transfer from air to sea", In: Bolin, B. (Ed.): Carbon cycle
is html
modelling  (SCOPE  16),  John  Wiley  &  Sons,  pp.  249-257,
http://iridl.ldeo.columbia.edu/SOURCES/.NOAA/.NCDC/.G
1981.
HCN/.v2/?bbox=bb%3A-161.488%3A16.360%3A-150.062%
3A23.051%3Abb
[40] M.  L.  Salby,  "Physics  of  the  Atmosphere  and  Climate",
Cambridge  University  Press,  Cambridge  2012.  (ISBN:
[27] NOAA, 2018:
978-0-521- 76718-7).
http://iridl.ldeo.columbia.edu/SOURCES/.NOAA/.NCDC/.ER
SST/.version2/.SST/index html
[41] D.  M.  Etheridge,  L.  P.  Steele,  R.  L.  Langenfelds,  R.  J.  Francey,
http://iridl.ldeo.columbia.edu/SOURCES/.NOAA/.NCDC/.ER
J.-M. Barnola, V. I. Morgan, "Natural and anthropogenic changes
SST/.version2/.SST/X/%28164W%29VALUES/T/%28Jan%2
in atmospheric CO2 over the last 1000 years from air in Antarctic
01938%29%28Dec%202009%29RANGEEDGES/Y/%2819N
ice and firn", J. Geophys. Res. 101, pp. 4115-4128, 1996. 
%29VALUES/datafiles.html
[42] Friedli H., H. Lötscher, H. Oeschger, U. Siegenthaler, B. Stauffer,
[28] O.  Humlum,  K.  Stordahl,  J.  E.  Solheim,  "The  phase  relation
1986. Ice core record of the 13C/12C ratio of atmospheric CO2 in
between atmospheric carbon dioxide and global temperature",
the past two centuries, Nature 324, pp. 237-238. 
Global and Planetary Change 100, pp. 51-69, 2013.
[43] H. Suess, "Radiocarbon Concentration in Modern Wood",
[29] M.  Richardson,  Comment  on  “The  phase  relation  between
Science 122, Issue 3166, pp. 415-417, 1955. DOI:
atmospheric carbon dioxide and global temperature” by Hum-
10.1126/science.122 3166.415-a
lum, Stordahl and Solheim, Global and Planetary Change 107,
pp. 226-228, 2013.
[44] J.  G.  Canadell   Le  Quéré,  C.,  Raupach,  M.  R.,  Field,  C.  B.,
Buitenhuis,  E.  T.,  Ciais,  P ,  Conway,  T.  J.,  Gillett,  N.  P.,
[30] D. L. Royer, R. A. Berner, I. P. Montañez, N. J. Tabor, D. J.
Houghton,  R.  A.,  and  Marland  G.,  "Contributions  to
Beerling, "CO
accelerating atmospheric CO
2 as a primary driver of Phanerozoic climate",
2 growth from economic activity,
GSA Today 14, no. 3, 2004.
carbon intensity, and efficiency of natural sinks", P. Natl. Acad.
ACT 1982
doi: 10.1130/1052-5173(2004)014<4:CAAPDO>2.0.CO;2.
USA, 
104(47), 
18866–18870, 
2007,
doi 10 1073/pnas 0702737104.
[31] Y. G. Zhang, M. Pagani, J. Henderiks, H. Ren, "A long history
of equatorial deep-water upwelling in the Pacific Ocean", Earth
[45] M. Gloor, J. L. S rmiento, and N. Gruber, "What can be learned
and Planetary Science Letters 467, pp. 1–9, 2017.
about  carbon  cycle  climate  feedbacks  from  the  CO2  airborne
http://dx.doi.org/10.1016/j.epsl.2017.03.016.
fraction"?  Atmos.  Chem.  Phys.,  10,  pp.  7739–7751,  2010.
https://www.atmos-chem-phys net/10/7739/2010/,
[32] T. Takahashi, S. C. Sutherland, R. Wanninkhof, C. Sweeney, R
doi:10.5194/acp-10-7739-2010.
A. Feely et al., "Climatological mean and decadal change in
surface ocean pCO
[46] A. P. Ballantyne, C. B. Alden, J. B. Miller, P. P. Tans, J. W. C.
2 and net sea-air CO2 flux over the global 
oceans", Deep-Sea Res. II, 56, pp. 554–577, 2009. 
White, "Increase in observed net carbon dioxide uptake by land
doi:10.1016/j.dsr2.2008.12.009. 
and  oceans  during  the  past  50  years",  Nature  488, pp. 70-73,
2012. doi:10.1038/nature11299
[33] N.  U.  Benson,  O.  O.  Osibanjo,  F.  E.  Asuquo,  W.  U.  Anake
"Observed trends of pCO
[47] J. Steele, "How NOAA and Bad Modeling Invented an Ocean
2 and air-sea CO2 fluxes in the North
Atlantic Ocean, Intern. J. Marine Science 4, pp. 1-7, 2014.
Acidification Icon", Part 2 – Bad Models, 2017.
https://wattsupwiththat.com/2017/03/02/how-noaa-and-bad-mo
[34] J.-S.  Lee,  "Monitoring  soil  respiration  using  an  automatic
deling-invented-an-ocean-acidification-icon-part-2-bad-models/
operating  chamber  in  a  Gwangneung  temperate  deciduous
forest", J. Ecology & Field Biology 34(4), pp. 411-423, 2011.
[48] W.  Evans,  B.  Hales,  P.  G.  Strut,  "Seasonal  cycle  of  surface
INFORMATION 
ocean pCO2 on the Oregon shelf", J. Geophys. Research 116,
[35] C. Huntingford, O. K. Atkin, A. Martinez de la Torre, L. M.
2011, DOI: 10.1029/2010JC006625.
Mercado, M. A. Heskel, A. B. Harp r, K. J. Bloomfield, O. S.
RELEASED UNDER THE 
O’Sullivan, P. B. Reich, K. R  Wythers, E. E. Butler, M. Chen,
[49] K. R. Arrigo, G. L. van Dijken, "Continued increases in Arctic
K. L. Griffin, P. Meir, M. G  Tjoelker, M. H. Turnbull, S. Sitch,
Ocean primary production", Progress in Oceanography 136,
A. Wiltshire, Y. Malhi, "Implications of improved
pp. 60-70, 2015, https://doi.org/10.1016/j.pocean.2015.05.002.
representations of plant respiration in a changing climate",
NATURE COMMUNICATIONS 8, 1602, 2017.
[50] D. K. Steinberg, M. W. Lomas, J. S. Cope, "Long-term increase
DOI: 10.1038/s41467-017-01774-z.
in  mesozooplankton  biomass  in  the  Sargasso  Sea:  Linkage  to
climate  and  implications  for  food  web  dynamics  and
[36] H. Harde, "Radia ion Transfer Calculations and Assessment of
biogeochemical  cycling",  Global  Biogeochemical  Cycle  26,
Global  Warming  by  CO
2012, DOI: 10.1029/2010GB004026.
2", 
International  Journal  of
Atmospheric Sciences, Volume 2017, Article ID 9251034, pp.
 OFFICIAL 
1-30, 2017. https://doi.org/10.1155/2017/9251034.
[51] K. M. Krumhardt, N. S. Lovenduski, N. M. Freeman, N. R.
Bates, "Apparent increase in coccolithophore abundance in the
[37] H.  Harde,  "Was  tragen  CO
subtropical North Atlantic from 1990 to 2014", Biogeosciences
2  und  die  Sonne  zur  globalen
Erwärmung  bei"?  12.  Internationale  EIKE  Klima-  und
13, pp. 1163-1177, 2016. doi:10.5194/bg-13-1163-2016,
Energiekonferenz  und  13th  International  Conference  on
http://www.biogeosciences net/13/1163/2016/.
Climate  Change  (ICCC-13),  München,  23.  u.  24.  November,
2018, https://youtu.be/ldrG4mn_KCs.
[52] F. P. Chavez, M. Messié, J. T. Pennington, "Marine Primary
Production in Relation to Climate Variability and Change",
[38] T.  B.  Coplen,  "Reporting  of  stable  hydrogen,  carbon  and
Annu. Rev. Mar. Sci. 3, pp. 227–260, 2011,
oxygen isotopic abundances", Pure and Applied Chemistry 66,
doi:10.1146/annurev.marine.010908.163917.
pp. 273-276, 1994.

159 
Hermann Harde: What Humans Contribute to Atmospheric CO2: Comparison of Carbon Cycle Models with Observations 
[53] U.  Riebesell,  K.  G.  Schulz,  R.  G.  J.  Bellerby,  M.  Botros,  P.
"Deep  ocean  nutrients  imply  large  latitudinal  variation  in 
Fritsche, M. Meyerhöfer, C. Neill, G. Nondal, A. Oschlies, J.
particle transfer efficiency", PNAS 113 no. 31, pp. 8606–8611, 
Wohlers, E. Zöllner, "Enhanced biological carbon consumption
2016, doi: 10.1073/pnas.1604414113. 
in  a  high  CO2  ocean",  Nature  450,  pp.  545-548,  2007,
doi:10.1038/nature06267.
[56] F. Abrantes, P. Cermeno, C. Lopes, O. Romero, L. Matos, J.
Van Iperen, M. Rufino, V. Magalhães, "Diatoms Si uptake
[54] A. M. P. McDonnell, K. O. Buesseler, "Variability in the
capacity drives carbon export in coastal upwelling systems",
average sinking velocity of marine particles", Limnology and
Biogeosciences 13, pp. 4099–4109, 2016,
Oceanography 55, pp. 2085–2096, 2010.
https://doi.org/10.5194/bg-13-4099-2016
DOI:10.4319/lo.2010.55.5.2085.
[57] M. Stuiver, H. A. Polach, "Discussion Reporting of 14C Data",
[55] T.  Weber,  J.  A.  Cram,  S.  W.  Leung,  T.  DeVries,  C.  Deutsch,
RADIOCARBON 19, No. 3, pp. 355-363, 1977.
ACT 1982
INFORMATION 
RELEASED UNDER THE 
 OFFICIAL 

GEOPHYSICAL RESEARCH LETTERS, VOL. 34, L01602, doi:10.1029/2006GL028492, 2007
On the decadal rates of sea level change during the twentieth century
S. J. Holgate1
Received 17 October 2006; accepted 21 November 2006; published 4 January 2007.
[1]
Nine long and nearly continuous sea level records were
problem that not all tide gauge records are of equivalent
chosen from around the world to explore rates of change in
quality. This can either be due to their location (being for
sea level for 1904 – 2003. These records were found to
example in an earthquake-prone region or an area of high
capture the variability found in a larger number of stations
glacial isostatic adjustment, GIA) or due to the quality of the
over the last half century studied previously. Extending the
instrumental record (being perhaps too discontinuous or
sea level record back over the entire century suggests that
lacking critical datum information to account for lo al
the high variability in the rates of sea level change observed
vertical land movements).
over the past 20 years were not particularly unusual. The
[5] As a result of these two problems, there are very few
rate of sea level change was found to be larger in the early
high quality, long tide gauge records in different regions
part of last century (2.03 ± 0.35 mm/yr 1904 – 1953),
suitable for calculating global mean sea level change. An
in comparison with the latter part (1.45 ± 0.34 mm/yr
alternative approach is to make use of regional composites
1954 – 2003). The highest decadal rate of rise occurred in
of shorter records as in HW04.
the decade centred on 1980 (5.31 mm/yr) with the lowest
[6] In order to test whether a few high quality records
rate of rise occurring in the decade centred on 1964
could provide similar information to the composites, nine
( 1.49 mm/yr). Over the entire century the mean rate of
tide gauge records were carefully selected from the database
ACT 1982
change was 1.74 ± 0.16 mm/yr. Citation: Holgate, S. J.
of the Permanent Service for Mean Sea Level (PSMSL,
(2007), On the decadal rates of sea level change during the
available at http://www.p l.ac.uk/psmsl) [Woodworth and
twentieth century, Geophys. Res. Lett., 34, L01602, doi:10.1029/
Player, 2003]: New York (1856 – 2003), Key West (1913 –
2006GL028492.
2003), San Diego (1906 – 2003), Balboa (1908 – 1996),
Honolulu (1905 – 2003), Cascais (1882 – 1993), Newlyn
(1915 – 2004) Trieste (1905 – 2004), and Auckland (1903 –
1.
Introduction
2000). The nine long records thus enable the study of
[2] In a previous paper, Holgate and Woodwo th [2004]
HW04 into variability of decadal rates of sea level change
(hereinafter referred to as HW04), rates of mean
global’’
to be extended over a much longer period. The locations of
sea level change (i.e., global coastal sea level change) were
these tide gauge stations are shown in Figure 1.
calculated from a large number of tide gauge records (177)
[7] These tide gauge stations are part of the Revised
for the period 1955 – 1998. HW04 found that the highest
Local Reference (RLR) data set of the PSMSL in which
and lowest rates of change in the 1955 – 1998 period
each time series is recorded relative to a consistent reference
occurred in the last 20 years of the record. In this paper it
level on the nearby land. Annual values in the RLR data set
is examined whether a few high quality tide gauge records
of the PSMSL are only calculated if there are at least
can replace the many used by HW04. On the basis of these
11 months of data and each month must have less than
high quality records the work of HW04 is then extended
15 missing days. Hence the tide gauge data presented here is
INFORMATION 
back to the early twentieth century to examine whether the
of the very highest quality available. All these records are
rates of sea level change experienced in recent decades are
almost continuous and are far away from regions with high
unusual.
rates of vertical land movement due to GIA or tectonics.
RELEASED UNDER THE 
[3] On a decadal timescale, the length scales of sea level
[8] Although most of these tide gauge records continue to
change are very large (O(1000) km) though not necessarily
the present, submissions of data to the PSMSL are often a
global. As a result, many tide gauges in a given region are
year or two in arrears and hence most of these sea level
highly correlated with each other. This paper demonstrates
records have data up until only 2003 or 2004. The current
that a few high quality records from around the world can
analysis begins in 1904 and ends in 2003 which ensures at
be used to examine large spatial-scale decadal variability as
least 70% completeness of the record in every decade.
well as many gauges from each region are able to.
[9] Following the method described in HW04, consecu-
tive, overlapping decadal mean rates were calculated for
each sea level record. The advantage of calculating decadal
 OFFICIAL 
2.
Method
rates in this way is that the tide gauge records can then be
[4] When it comes to calculating long term global sea
combined into a single mean sea level time series, despite
level means from tide gauge data, there are a number of
the different gauges having different datums. Furthermore,
problems. Firstly there is a bias in the distribution of tide
decadal rates remove any minor data discontinuities and
gauges towards certain regions, notably Northern Europe
introduce an element of smoothing. The rates of change at
and North America [Douglas, 1991]. Secondly there is the
each station are corrected for GIA using the ICE-4G model
of Peltier [2001] and for inverse barometer effects using the
HadSLP2 air pressure data set [Allan and Ansell, 2006].
1Proudman Oceanographic Laboratory, Liverpool, UK.
[10] The standard error of a sea level trend estimate,
based on the assumption that each annual mean is inde-
Published in 2007 by the American Geophysical Union.
L01602
1 of 4


ACT 1982
INFORMATION 
RELEASED UNDER THE 
 OFFICIAL 

L01602
HOLGATE: THE 20TH CENTURY SEA LEVEL CHANGE
L01602
between the global mean and Trieste is 0.49 in comparison
with the difference between the global mean and New York
(the highest individual rate) which is 0.62. It would there-
fore seem that Trieste no more biases the mean low than
New York biases the mean high. Nevertheless, excluding
Trieste from the results would slightly increase the global
mean from 1.74 to 1.80 mm/yr.
[20] Although the mean rate of change of global mean sea
level is found to be greater in the first half of the twentieth
century, the two rates are consistent with being the same at
the 95% confidence level, given their individual standard
errors. However, a greater rate of rise in the early part of the
record is consistent with previous analyses of tide gauge
records which suggested a general deceleration in sea level
rise during the 20th century [Woodworth, 1990; Douglas,
1992; Jevrejeva et al., 2006]. A twentieth century deceler-
ation is consistent with the work of Church and White
[2006] who, although finding evidence for a post-1870
acceleration based on an EOF reconstruction of global sea
level, found that much of the overall acceleration occurred
in the first half of the 20th century. Church and White
Figure 3. Comparison of the decadal rates of sea level
[2006] sugges ed that the greater rate of sea level rise
change for each of the nine records. All rates are corrected
observed in the first half of last century was due to reduced
for glacial isostatic adjustment and inverse barometer
ACT 1982
volcanic emissions (and hence also lower variability in sea
effects.
level) during the 1930s to 1960s. This idea is supported by
results from the HadCM3 model which suggest that the
Cascais (1.85 ± 0.37 mm/yr). The smallest changes in sea
simulated global mean sea level did not accelerate through
level are seen in Trieste (1.25 ± 0.23 mm/yr) and Newlyn
the twentieth century due to the offsetting of anthropogenic
(1.46 ± 0.30 mm/yr).
warming by reduced natural forcing [Gregory et al., 2006].
[16] San Diego has the highest correlation with the global
[21] The decadal rates of sea level change shown in
mean rates (r = 0.62) over the 1904 – 2003 period, followed
Figure 2 are qualitatively similar to the corresponding rates
by Honolulu (r = 0.58), New York (r = 0.56), Balboa (r =
in Figure 2 of Church and White [2006], with the exception
0.55) and Trieste (r = 0.42). Cascais and Auckland have
of the period 1930 – 1940 which shows lower variability in
insignificant correlations at the 95% confidence level while
the work of Church and White [2006]. The variability in the
the correlations with Newlyn (r = 0.29) and Key West (r =
second half of the century is also similar to that found by
0.25) are significant but low.
4.
Discussion
[17] The nine stations selected here as high quality
records capture the mean decadal rates of change described
INFORMATION 
by the larger set of stations used in HW04 and also have a
similar global mean rate over the common period of the two
RELEASED UNDER THE 
analyses (1953 – 1997). This provides confidence that the
nine station set can be used to study decadal rates of global
mean sea level change throughout the twentieth century.
[18] All the stations in this study show a significant
increase in sea level over the period 1904 – 2003 with an
average increase of 174 mm during that time (Figure 4).
This mean rate of 1.74 mm/yr is at the upper end of the
range of estimates for the 20th century in the Intergovern-
mental Panel on Climate Change, Third Assessment Report
 OFFICIAL 
(IPCC TAR) [Church et al., 2001], and consistent with
other recent estimates [Holgate and Woodworth, 2004;
Church and White, 2006].
[19] The rates for individual stations are consistent with
those published by other authors [Douglas, 2001; Peltier,
2001; Hannah, 1990]. As has been noted previously
[Woodworth, 1990], the rates for northern European tide
gauges are consistently lower than the global mean. Trieste,
Figure 4. The mean sea level record from the nine tide
along with other Mediterranean tide gauge stations, has
gauges over the period 1904 – 2003 based on the decadal
shown a much lower rate of increase since 1960 [Douglas,
trend values for 1907 – 1999. The sea level curve here is the
1997; Tsimplis and Baker, 2000]. However, the difference
integral of the rates presented in Figure 2.
3 of 4

L01602
HOLGATE: THE 20TH CENTURY SEA LEVEL CHANGE
L01602
Chambers et al. [2002] though the lower number of gauges
mate Change, edited by J. T. Houghton et al., chap. 11, pp. 639 694,
Cambridge Univ. Press, New York.
in the present study results in a greater level of variance.
Douglas, B. C. (1991), Global sea level rise, J. Geophys. Res., 96, 6981
6992.
Douglas, B. C. (1992), Global sea level acceleration, J. Geophys. Res., 97,
5.
Summary and Conclusions
12,699 12,706.
[
Douglas, B. C. (1997), Global sea rise: A redetermination, Surv. Geophys.,
22]
Based on a selection of nine long, high quality tide
18, 279 292.
gauge records, the mean rate of sea level rise over the period
Douglas, B. C. (2001), Sea level change in the era of the recording tide
1904 – 2003 was found to be 1.74 ± 0.16 mm/yr after
gauge, in Sea Level Rise: History and Consequences, Int. Geophys. Ser.,
correction for GIA using the ICE-4G model [Peltier,
vol. 75, edited by B. C. Douglas, M. S. Kearney, and S. P. Leatherman,
chap. 3, pp. 37 64, Elsevier, New York.
2001] and for inverse barometer effects using HadSLP2
Gregory, J., J. Lowe, and S. Tett (2006), Simulated global-mean sea-level
[Allan and Ansell, 2006]. The mean rate of rise was greater
changes over the last half-millenium, J. Clim., 19, 4576 4591.
in the first half of this period than the latter half, though the
Hannah, J. (1990), Analysis of mean sea level data from New Zealand for
the period 1899 1988, J. Geophys. Res., 95, 12,399 12,405.
difference in rates was not found to be significant. The use
Holgate, S. J., and P. L. Woodworth (2004), Evidence for enhanced coastal
of a reduced number of high quality sea level records was
sea level rise during the 1990s, Geophys. Res. Lett., 31, L07305,
found to be as suitable in this type of analysis as using a
doi:10.1029/2004GL019626.
larger number of regionally averaged gauges.
Jevrejeva, S., A. Grinsted, J. C. Moore, and S. Holgate (2006 , Nonlinear
trends and multi-year cycles in sea level trends, J. Geophys. Res., 111,
[23] Finally, in extending the work of HW04 to cover
C09012, doi:10.1029/2005JC003229.
the whole century, it is found that the high decadal rates of
Maul, G. A., and D. M. Martin (1993), Sea level rise at Key West, Florida,
change in global mean sea level observed during the last
1846 1992: America’s longest ins rument record?, Geophys. Res. Lett.,
20, 1955 1958.
20 years of the record were not particularly unusual in the
Nerem, R. S., and G T Mitchum (2002), Estimates of vertical crustal
longer term context.
motion derived from differences of TOPEX/POSEIDON and tide gauge
sea level measurements, Geophys. Res Lett , 29(19), 1934, doi:10.1029/
2002GL015037.
[24] Acknowledgments. I’d like to thank Phil Woodworth, Simon
Peltier, W. (2001), Global glacial isostatic adjustment and modern instru-
Williams, and Svetlana Jevrejeva for discussion and comments which have
ACT 1982
mental records of relative s a level history, in Sea Level Rise: History and
helped to improve this paper.
Consequences, Int. Geophys. Ser., vol. 75, edited by B. C. Douglas, M. S.
Kearney, a d S. P. Leath rman, chap. 4, pp. 65 95, Elsevier, New York.
References
Tsimplis, M. N., and T. F Baker (2000), Sea level drop in the Mediterra-
nean Sea: An indicator of deep water salinity and temperature changes?,
Allan, R., and T. Ansell (2006), A new globally complete monthly histor-
G
phys. Res. Lett., 27, 1731 1734.
ical mean sea level pressure data set (HadSLP2): 1850 2004, J. Clim, in
Woodworth P (1990), A search for accelerations in records of European
press.
mean sea level Int. J. Climatol., 10, 129 143.
Chambers, D. P., C. A. Mehlhaff, T. J. Urban, D. Fujii, and R. S. Ner m
Woodworth, P., and R. Player (2003), The Permanent Service for Mean Sea
(2002), Low-frequency variations in global mean sea level: 1950 2000,
Level: An update to the 21st century, J. Coastal Res., 19(2), 287 295.
J. Geophys. Res., 107(C4), 3026, doi:10.1029/2001JC001089.
World Meteorological Organization (1966) Report of a working group on
Church, J. A., and N. J. White (2006), A 20th century acceleration in
the commission for climatology, Tech. Rep. 79, 79 pp., World Meteorol.
global sea level rise, Geophys. Res. Lett., 33, L01602, doi:10.1029/
Organ., Geneva, Switzerland.
2005GL024826.
Church, J. A., J. Gregory, P. Huybrechts, M. Kuhn, K. Lambeck, M. Nhuan,
D. Qin, and P. Woodworth (2001), Changes in sea level, in Climate
Change 2001: The Scientific Basis: Contribution of Working Group
S. J. Holgate, Proudman Oceanographic Laboratory, Joseph Proudman
to the Third Assessment Report of the Intergovernmental Panel on Cli-
Building, 6 Brownlow Street, Liverpool L3 5DA, UK. ([email address])
INFORMATION 
RELEASED UNDER THE 
 OFFICIAL 
4 of 4


ACT 1982
INFORMATION 
RELEASED UNDER THE 
 OFFICIAL 

Citation: N kolov N, Zeller K (2017) New Insights on the Physical Nature of the Atmospheric Greenhouse Effect Deduced from an Empirical Planetary 
Temperature Model. Environ Pollut Climate Change 1: 112. 
Page 2 of 22
GCMs, we decided to try an empirical approach not constrained by a 
The simplifying power of DA in model development stems from the 
particular physical theory. An important reason for this was the fact that  Buckingham Pi Theorem [27], which states that a problem involving n 
current process-oriented climate models rely on numerous theoretical  dimensioned x  variables, i.e.
i
assumptions while utilizing planet-specific parameterizations of key 
processes such as vertical convection and cloud nucleation in order 
x , x , , x =
n
0
1
2
)
to simulate the surface thermal regime over a range of planetary  can be reformulated into a simpler relationship of (n-m) dimensionless 
environments [15]. These empirical parameterizations oftentimes  π products derived from x , i.e.
i 
i
depend on detailed observations that are not typical y available for 
ϕ(π π , …. ,π )  =  0
planetary bodies other than Earth.  Hence, our goal was to develop 
1
2
n-m
a simple yet robust planetary temperature model of high predictive  where  m is the number of fundamental dimensions comprising the 
power that does not require case-specific parameter adjustments while  original variables. This theorem determines the number of non-
successful y describing the observed range of planetary temperatures  dimensional π  variables to be found in a set of products, but it does not 
i
across the Solar System. 
prescribe the number of sets that could be generated from the original 
variables defining a particular problem. In other words, there might be, 
Methods and Data
and oftentimes is more than one set of (n-m) dimensionless products to 
analyze. DA provides an objective method for constructing the sets of 
In our model development we employed a ‘top-down’ empirical  π variables employing simultaneous equations solved via either matrix 
approach based on Dimensional Analysis (DA) of observed data 
i
inversion or substit tion [22]. 
from our Solar System. We chose DA as an analytic tool because of 
its ubiquitous past successes in solving complex problems of physics, 
The second step of DA (after the construction of dimensionless 
engineering, mathematical biology, and biophysics [16-21]. To our  products) is to search for a functional relationship between the π  i
knowledge DA has not previously been applied to constructing  variables of e ch set using regression analysis. DA does not disclose 
ACT 1982
predictive models of macro-level properties such as the average global  the best function capable of describing the empirical data. It is the 
temperature of a planet; thus, the following overview of this technique  investigator’s resp nsibility to identify a suitable regression model 
is warranted.
based on prior knowledge of the phenomenon and a general expertise 
in the subject area  DA only guarantees that the final model (whatever 
Dimensional analysis background
its functional form) will be dimensional y homogeneous, hence it may 
DA is a method for extracting physical y meaningful relationships  qualify as a physical y meaningful relationship provided that  it (a) is 
from empirical data [22-24]. The goal of DA is to restructure   set of  not b sed  n a simple polynomial fit; (b) has a small standard error; 
original variables deemed critical to describing a physical phenomenon  (c) displays high predictive skill over a broad range of input data; and
into a smaller set of independent dimensionless products that may be  (d) is statistical y robust. The regression coefficients of the final model
will also be dimensionless, and may reveal true constants of Nature by
combined into a dimensional y homogeneous model with predic ive  virtue of being independent of the units utilized to measure the forcing
power. Dimensional homogeneity is a prerequisite for any robust  variables.
physical relationship such as natural laws. DA distinguishes  etween 
measurement units  and  physical dimensions. For example, mass is a  Selection of model variables
physical dimension that can be measured in gram, pound, metric ton 
A planet’s GMAT depends on many factors. In this study, we focused 
etc.; time is another dimension measurable in seconds, hours, years,  on drivers that are remotely measurable and/or theoretical y estimable. 
etc. While the physical dimension of a variable does not change, the  Based on the current state of knowledge we identified seven physical 
INFORMATION 
units quantifying that variable may vary depending on the adopted  variables of potential relevance to the global surface temperature: 1) top-
measurement system. 
of-the-atmosphere (TOA) solar irradiance (S); 2) mean planetary surface 
RELEASED UNDER THE 
Many physical variables and constant  can be described in terms of four  temperature in the absence of atmospheric greenhouse effect, hereto 
fundamental dimensions, i.e. mass [M], length [L], time [T], and absolute  called a reference temperature (); 3) near-surface partial pressure 
r
temperature [Θ]. For example, an energy flux commonly measured in W  of atmospheric greenhouse gases (); 4) near-surface mass density 
gh
m
of atmospheric greenhouse gases (ρ ); 5) total surface atmospheric 
-2 has a physical dimension [M T ] since 1 W m-2 = 1 J s-1 m-2 = 1 (kg m2
gh
s
pressure (P); 6) total surface atmospheric density (ρ); and 7) minimum 
-2) s-1 m-2 = kg s-3. Pressure may be reported in units of Pascal, bar, atm.,
air pressure required for the existence of a liquid solvent at the surface, 
PSI or Torr, but its physical dimension is always [M L-1 T-2] because 1 Pa
hereto called a reference pressure (). Table 1 lists the above variables 
= 1 N m-2 = 1 (kg m s- ) m 2 = 1 kg m-1 s-2. Thinking in terms of physical
r
along with their SI units and physical dimensions. Note that, in order to 
dimensions rather than measurement units fosters a deeper understanding  simplify the derivation of dimensionless products, pressure and density 
of the underlying physical reality. For instance, a comparison between
are represented in Table 1 by the generic variables  and ρ , respectively. 
 OFFICIAL 
the physical dimensions of energy flux and pressure reveals that a flux is
x
x
As explained below, the regression analysis following the construction 
simply the product of pressure and the speed of moving particles [L T-1],
of  π  variables explicitly distinguished between models involving 
i.e. [M T
i
-3] = [M L-1 T-2] [L T-1]. Thus, a radiative flux  (W m-2) can be
R
partial pressure/density of greenhouse gases and those employing total 
expressed in terms of photon pressure  (Pa) and the speed of light c (m
ph
atmospheric pressure/density at the surface. The planetary Bond albedo 
s-1) as  = c P . Since c is constant within a medium, varying the intensity
(α ) was omitted as a forcing variable in our DA despite its known effect 
R
ph
p
of electromagnetic radiation in a given medium effectively means altering
on the surface energy budget, because it is already dimensionless and 
the pressure of photons. Thus, the solar radiation reaching Earth’s upper
also partakes in the calculation of reference temperatures discussed 
atmosphere exerts a pressure (force) of sufficient magnitude to perturb the  below.
orbits of communication satellites over time [25,26]. 
Appendix A details the procedure employed to construct the π  i
variables. DA yielded two sets of π  products, each one consisting of two 
i
Environ Pollut Climate Change, an open access journal 
Volume 1 • Issue 2 • 1000112

Citation: N kolov N, Zeller K (2017) New Insights on the Physical Nature of the Atmospheric Greenhouse Effect Deduced from an Empirical Planetary 
Temperature Model. Environ Pollut Climate Change 1: 112. 
Page 3 of 22
Planetary Variable
Symbol
SI Units
Physical Dimension
Global mean annual near-surface temperature (GMAT), the dependent variable 
T
K
[Θ]
s
Stellar irradiance (average shortwave flux incident on a plane perpendicular to the stellar rays at the top of a planet s 
atmosphere)
S
W m-2
[M T-3]
Reference temperature (the planet’s mean surface temperature in the absence of an atmosphere or an atmospheric 
greenhouse effect)
T
K
[Θ]
r
Average near-surface gas pressure representing either partial pressure of greenhouse gases or total atmospheric 
pressure
P
Pa
[M L-1 T-2]
x
Average near-surface gas density representing either greenhouse-gas density or total atmospheric density
ρ
kg m-3
[M L-3]
x
Reference pressure (the minimum atmospheric pressure required a liquid solvent to exists at the surface)
P
Pa
[M L-1 T-2]
r
Table 1: Variables employed in the Dimensional Analysis aimed at deriving a general planetary temperature model. The variables are comprised of 4 fundamental physical 
dimensions: mass [M], length [L], time [T] and absolute temperature [Θ]. 
dimensionless variables, i.e.
effect, both based on the SB radiation law. The first and most popular 
3
T
P
approach uses the planet’s global energy budget to calculate a single 
s
π =
        
x
π =
1
;
2
2
radiating equilibrium temperature  (also known as an effective 
T
ρ    S
e
r
x
and
emission temperature) from  he average absorbed solar flux [6,9,28], 
i.e.
 
T
P
s
π =
            
x
π =
.
1
;
2

− α 
T
P
 
( 1
) 0 25
r
r

e


(3)
This implies an investigation of two types of dimensional y homogeneous 
4  
εσ


functions (relationships): 
Here,  S is the solar irradiance (W m-2) defined as the TOA 
ACT 1982
3
T
 
s
x
= ƒ 
  (1)
shortwave flux incident on a plane perpendicular to the incoming rays, 
2  
T
ρ  
 
S
r
 x

α  is  he planetary Bond albedo (decimal fraction), ε is the planet’s 
and
p
LW emissivity (typical y 0.9 ≤  ε <1.0; in this study we assume ε = 0.98 
T
 
b sed on lun r regolith measurements reported by Vasavada et al. [29], 
s
x
=
(2)

         
and σ = 5 6704 × 10-8 W m-2 K-4 is the SB constant. The term S(1-α  )⁄4 
T
P
p
r
 
represents a global y averaged shortwave flux absorbed by the planet-
Note that π  = T /T  occurs as a dependent variable in both relationships, 
1
s
r
atmosphere system. The rationale behind Eq. (3) is that the TOA energy 
since it contains the sought temperature . Upon replacing the generic 
s
b l nce presumably defines a baseline temperature at a certain height 
pressure/density variables P   and  ρ  in functions (1) and (2) with 
x
x
in  he free atmosphere (around 5 km for Earth), which is related to the 
either partial pressure/density of greenhouse gases (and  ) or total 
gh 
gh
planet’s mean surface temperature via the infrared optical depth of the 
atmospheric pressure/density (P and ρ), one arr ves at six prospective 
atmosphere [9,10]. Equation (3) was introduced to planetary science 
regression models. Further, as explained below, we employed two 
in the early 1960s [30,31] and has been widely utilized ever since to 
distinct kinds of reference temperature computed from different  calculate the average surface temperatures of airless (or nearly airless) 
formulas, i.e. an effective radiating equilibrium temperature () and 
e
bodies such as Mercury, Moon and Mars [32] as well as to quantify 
a mean ‘no-atmosphere’ spherical surface temperature (). This 
na
the strength of the greenhouse effect of planetary atmospheres [2-
doubled the π  instances in the regression analysis bringing the total 
i
4,6,9,28]. However, Volokin and ReLlez [1] showed that, due to Hölder’s 
number of potential models for investigation to twelv
inequality between integrals [33],  is a non-physical temperature for 
INFORMATION  e
Reference temperatures and reference pressure
spheres and lacks a meaningful relationship to the planet’s s
A reference temperature () characterizes the average thermal 
The second method attempts to estimate the average surface 
RELEASED UNDER THE 
r
environment at the surface of a planetary body in the absence of 
temperature of a planet () in the complete absence of an atmosphere 
na
atmospheric greenhouse effect; hence,  is different for each body and 
using an explicit spatial integration of the SB law over a sphere. Instead 
r
depends on solar irradiance and surface albedo. The purpose of  is 
of calculating a single bulk temperature from the average absorbed 
r
to provide a baseline for quantifying the thermal effect of planetary 
shortwave flux as done in Eq. (3), this alternative approach first 
atmospheres. Indeed, the T /T  ratio produced by DA can physical y be 
computes the equilibrium temperature at every point on the surface of 
s
r
interpreted as a Relative Atmospheric Thermal Enhancement (RATE) 
an airless planet from the local absorbed shortwave flux using the SB 
ideal y expected to be equal to or greater than 1.0. Expressing the 
relation, and then spherical y integrates the resulting temperature field 
thermal effect of a planetary atmosphere as a non-dimensional quotient 
to produce a global temperature mean. While algorithmical y opposite 
instead of an absolute temperature difference (as done in the past) 
to Eq. (3), this method mimics well the procedure for calculating Earth’s 
 OFFICIAL 
allows for an unbiased comparison of the greenhouse effects of celestial 
global temperature as an area-weighted average of surface observations.
bodies orbiting at different distances from the Sun. This is because the 
Rubincam [34] proposed an analytic solution to the spherical 
absolute strength of the greenhouse effect (measured in K) depends on 
integration of the SB law (his Eq. 15) assuming no heat storage by the 
both solar insolation and atmospheric properties, while RATE being 
regolith and zero thermal inertia of the ground. Volokin and ReLlez 
a radiation-normalized quantity is expected to only be a function of a 
[1] improved upon Rubincam’s formulation by deriving a closed-form
planet’s atmospheric environment. To our knowledge, RATE has not 
integral expression that explicitly accounts for the effect of subterranean 
previously been employed to measure the thermal effect of planetary 
heat storage, cosmic microwave background radiation (CMBR) and
atmospheres. 
geothermal heating on the average global surface temperature of
Two methods have been proposed thus far for estimating the 
airless bodies. The complete form of their analytic Spherical Airless-
average surface temperature of a planetary body without the greenhouse 
Temperature (SAT) model reads:
Environ Pollut Climate Change, an open access journal 
Volume 1 • Issue 2 • 1000112

Citation: N kolov N, Zeller K (2017) New Insights on the Physical Nature of the Atmospheric Greenhouse Effect Deduced from an Empirical Planetary 
Temperature Model. Environ Pollut Climate Change 1: 112. 
Page 4 of 22
 (1

 −η
can only exists in a solid/vapor phase and not in a liquid form. The results 
  S  
− α +  − R

)
(1 e)
54
c

C g)5/4

+ 
of our analysis are not sensitive to the particular choice of a reference-
2 
(1− η  S − α εσ

)
(1
 
) (
)1/4
 = 

           (4a)
pressure value; hence, the selection of P  is a matter of convention.
na
r
5 

 0.754η    S − α +  − R


e
(1
 
)
5/4
c

C g)5/4


Regression analysis

0.754η    S − α εσ

e
(1
 
) (
)1/4

Finding the best function to describe the observed variation of 
where  α  is the effective shortwave albedo of the surface, η  is the 
e
e
GMAT among celestial bodies requires that the π  variables generated 
effective ground heat storage coefficient in a vacuum, R  = σ 2.725
i
4 = 
c
by DA be subjected to regression analyses. As explained in Appendix A, 
3.13 × 10-6 W m-2 is the CMBR [35], and R  is the spatial y averaged 
g
twelve pairs of π  variables hereto called Models were investigated. In 
geothermal flux (W m
i
-2) emanating from the subsurface. The heat 
order to ease the curve fitting and simplify the visualization of results, 
storage term η  is defined as a fraction of the absorbed shortwave flux 
e
we utilized natural logarithms of the constructed π  variable  rather than 
conducted into the subsurface during daylight hour and subsequently 
i
their absolute values, i.e. we modeled the relationship ln (π ) = f (ln(π ))
released as heat at night. 
1
2
instead of π  = f(π ). In doing so we focused on monotonic functions 
1
2
Since the effect of CMBR on T  is negligible for > 0.15 W m-2 [1] 
of conservative shapes such as exponential, sigmoidal, hyperbolic, 
na 
and the geothermal contribution to surface temperatures is insignificant 
and logarithmic, for their fitting coefficients might be interpretable in 
for most planetary bodies, one can simplify Eq. (4a) by substituting  = 
physical y meaningful terms. A key advantage of this type of functions 
c
 = 0 This produces:
(provided the existence of a good fit, of course) is that they also tend 
g
to yield reliable results outside the data range used to determine their 
2  ( 1−α 
)
0.25
coefficients. We specifical y avoided non-monotonic functions such as 
=



−η
+
η

na
(1 )0.25
0.25
0.932 e
 
                  
5
ε  σ




                                           (4b)
polynomials because of thei   bility to accurately fit almost any dataset 
where 0.932 = 0.7540.25. The complete formula (4a) must only be used if 
given a sufficiently large number of regression coefficients while at the 
ACT 1982
S ≤ 0.15 W m-2 and/or the magnitude of  is significantly greater than 
same time showing poor predictive skil s beyond the calibration data 
g
zero. For comparison, in the Solar System, the threshold S ≤ 0.15 W m-2 
range  Due to their highly flexible shape, polynomials can easily fit 
is encountered beyond 95 astronomical unis (AU) in the region of the 
random noise in a d taset, an outcome we particularly tried to avoid.
inner Oort cloud. Volokin and ReLlez [1] verified Equations (4a) and 
The following four-parameter exponential-growth function was 
(4b) against Moon temperature data provided by the NASA Diviner 
found to  est m et our criteria:
Lunar Radiometer Experiment [29,36]. These authors also showed that 
accounting for the subterranean heat storage (η ) markedly improves 
a exp (b x) + c exp ( x)  
(5)
e
the physical realism and accuracy of the SAT model compared to the 
where x = ln (π )  and y = ln (π ) are the independent and dependent 
original formulation by Rubincam [34].
2
1
variable respectively while a, b, c and d are regression coefficients.  This 
The conceptual difference between Equations (3) and (4b) is tha  Τ  
function has a rigid shape that can only describe specific exponential 
e
represents the equilibrium temperature of a blackbody d sk orthogonal y 
patterns found in our data. Equation (5) was fitted to each one of the 
il uminated by shortwave radiation with an in ensity equal to the average 
12 planetary data sets of logarithmic π  pairs suggested by DA using the 
i
solar flux absorbed by a sphere having a Bond albedo α , while Τ  is the 
standard method of least squares. The skil s of the resulting regression 
p
na
area-weighted average temperature of a thermal y heterogen ous airless 
models were evaluated via three statistical criteria: coefficient of 
sphere [1,37]. In other words, for spherical objects  Τ  is an abstract 
determination (R2), adjusted R2, and standard error of the estimate (σ ) 
e
est
mathematical temperature, while T  is the average kinetic temperature 
[39,40]. All calculations were performed with SigmaPlotTM 13 graphing 
na 
of an airless surface. Due to Hölder’s inequality between integrals, one 
and analysis software.
INFORMATION 
always finds Τ >> Τ when us ng equivalent values of stel ar irradiance 

na 
Planetary data 
and surface albedo in Equations (3) and (4b) [1]
To ensure proper application of th
RELEASED UNDER THE e DA methodology we compiled a 
To calculate the T  temperatures for planetary bodies with tangible 
na 
dataset of diverse planetary environments in the Solar System using the 
atmospheres, we assumed that the airless equivalents of such objects 
best information available. Celestial bodies were selected for the analysis 
would be covered with a regolith of similar optical and thermo-physical 
based on three criteria: (a) presence of a solid surface; (b) availability 
properties as the Moon surface. This is based on the premise that, in 
of reliable data on near-surface temperature, atmospheric composition, 
the absence of a protective a mosphere, the open cosmic environment 
and total air pressure/density preferably from direct observations; and 
would erode and pulverize exposed surfaces of rocky planets over time 
(c) representation of a broad range of physical environments defined
in a similar manner [1]. Also, properties of the Moon surface are the 
in terms of TOA solar irradiance and atmospheric properties. This
best studied ones  mong all airless bodies in the Solar System. Hence, 
resulted in the selection of three planets: Venus, Earth, and Mars; and
one could further simplify Eq. (4b) by combining the albedo, the heat 
three natural satellites: Moon of Earth, Titan of Saturn, and Triton of
 OFFICIAL 
storage fraction and the emissivity parameter into a single constant 
Neptune.
using applicable values for the Moon, i.e. α  = 0.132, η  = 0.00971 and ε 
e
  e
= 0.98 [1,29]. This produces: 
Each celestial body was described by nine parameters shown in 
Table 2 with data sources listed in Table 3. In an effort to minimize 
0.25
=
S
           (4c)
the effect of unforced (internal) climate variability on the derivation 
na
32.44 
Equation (4c) was employed to estimate the ‘no-atmosphere’ reference 
of our temperature model, we tried to assemble a dataset of means 
temperatures of all planetary bodies participating in our analysis and 
representing an observational period of 30 years, i.e. from 1981 to 2010. 
discussed below. 
Thus, Voyager measurements of Titan from the early 1980s suggested 
an average surface temperature of 94 ± 0.7 K [41]. Subsequent 
For a reference pressure, we used the gas-liquid-solid triple point of 
observations by the Cassini mission between 2005 and 2010 indicated 
water, i.e.  = 611.73 Pa [38] defining a baric threshold, below which water 
r
a mean global temperature of 93.4 ± 0.6 K for that moon [42,43]. Since 
Environ Pollut Climate Change, an open access journal 
Volume 1 • Issue 2 • 1000112

Citation: N kolov N, Zeller K (2017) New Insights on the Physical Nature of the Atmospheric Greenhouse Effect Deduced from an Empirical Planetary 
Temperature Model. Environ Pollut Climate Change 1: 112. 
Page 5 of 22
Parameter
Venus
Earth
Moon
Mars
Titan
Triton
Average distance to the Sun,  (AU) 
0.7233
1.0
1.0
1.5237
9.582
30.07
au
Average TOA solar irradiance, S (W m-2)
2,601.3
1,360.9
1,360.9
586.2
14.8
1.5
Bond albedo, α  (decimal fraction)
0.900
0.294
0.136
0.235
0.265
0.650
p
Average absorbed shortwave radiation, S  = S(1-α )/4 (W m-2)
65.0
240.2
294.0
112.1
2.72
0.13
a
p
Global average surface atmospheric pressure, P (Pa)
9,300,000.0 ± 
100,000
98,550.0 ± 6.5
2.96 × 10-10 ± 
10
685.4 ± 14.2 146,700.0 ± 100
4.0 ± 1.2
-10
Global average surface atmospheric density, ρ (kg m
0.019 ± 3.2 × 
-3) 
65 868 ± 0.44
1.193 ± 0.002
2.81 × 10-15 ± 
9.4 × 10
5.161 ± 0.03
3.45 × 10-4 ± 9.2 
-15
10-4
× 10-5
77.89 N
26.7 
95.32 CO
4He
2
2
96.5 CO
20.89 O
26.7 
2.70 N
20Ne
2
99.91 N
2
2
2
Chemical composition of the lower atmosphere (% of volume)
3.48 N
0.932 Ar
23.3 H
1.60 Ar
95.1 N2
0 060 CO
2
2
0.02 SO
0.248 H O
20.0 
0.13 O
4.9 CH
40Ar
2
4
0.024 CH
2
2
4
0.040 CO
3.3 
0.08 CO
22Ne
2
0.021 H O
2
Molar mass of the lower atmosphere, (kg mol-1)
0.0434
0.0289
0.0156
0.0434
0.0274
0.0280
GMAT,  (K)
737.0 ± 3.0
287.4 ± 0.5
197.35 ± 0.9
190 56 ± 0.7
93.7 ± 0.6
39.0 ± 1 0
s
Table 2: Planetary data set used in the Dimensional Analysis compiled from sources listed in Table 3. The estimation of Mars’ GMAT and the average surface atmospheric 
pressure are discussed in Appendix B. See text for details about the computational methods employed for some parameters. 
Planetary Body
Information Sources
temperatures  bove 210 K can only occur on Mars during summertime. 
Venus
[32,44-48]
Hence, all  uch v lues must be significantly higher than the actual mean 
ACT 1982
Earth
[12,13,32,49-55]
annual temp rature at any M rtian latitude. This is also supported by 
results from a 3-D global circulation model of the Red Planet obtained 
Moon
 [1,29,32,48,56-59] 
by Fenton et al. [82]. The surface atmospheric pressure on Mars varies 
Mars
[32,48,60-63], Appendix B
appreciably with season and location. Its global average value has 
Titan
[32,41-43,64-72]
p eviously been reported between 600 Pa and 700 Pa [6,32,78,80,83,84], 
Triton
[48,73-75]
a range that was too broad for the target precision of our study. Hence 
Table 3: Literature sources of the planetary data presented in Table 2.
our decision to calculate new annual global means of near-surface 
Saturn’s orbital period equals 29.45 Earth years, we averaged th  above 
temperature and air pressure for Mars via a thorough analysis of available 
global temperature values to arrive at 93.7 ± 0.6 K as an estimate of 
dat  from remote-sensing and in-situ observations. Appendix B details 
Titan’s 30-year GMAT. Similarly, data gathered in the late 1970s by the 
our computational procedure with the results presented in Table 2. It is 
Viking Landers on Mars were combined with more recent Curiosity-
noteworthy that our independent estimate of Mars’ GMAT (190.56 ± 
Rover surface measurements and 1999-2005 remote observations by 
0.7 K), while significantly lower than values quoted in recent years, is in 
the Mars Global Surveyor (MGS) spacecraft to derive representative 
perfect agreement with spherical y integrated brightness temperatures 
estimates of GMAT and atmospheric  urface pressure for the Red 
of the Red Planet derived from remote microwave measurements in the 
Planet. Some parameter values reported in the literature did not meet 
late 1960s and early 1970s [85-87]. 
our criteria for global representativeness  nd/or physical plausibility 
Moon’s GMAT was also not readily extractable from the published 
and were recalculated using available  bservations a  described below.
literature. Although lunar temperatures have been measured for 
INFORMATION 
The mean solar irradiances of all bodies were calculated as S = S r -2 
more than 50 years both remotely and in situ [36] most studies focus 
E  au
where  is the body’s average distance (semi major axis) to the Sun 
on observed temperature extremes across the lunar surface [56] and 
au
(AU) and  = 1,360.9 W m-2 is the Earth’s new lower irradiance at 1 AU 
rarely discuss the Moon’s average global temperature. Current GMAT 
RELEASED UNDER THE 
E
according to recent satellite observations reported by Kopp and Lean 
estimates for the Moon cluster around two narrow ranges: 250–255 
[49]. Due to a design flaw in earlier spectrometers, the solar irradiance 
K and 269–271 K [32]. A careful examination of the published data 
at Earth’s distance has been overestimated by ≈ 5 W m-2 prior to 2003 
reveals that the 250–255 K range is based on subterranean heat-flow 
[49]. Consequently, our calculations yielded slightly lower irradiances 
measurements conducted at depths between 80 and 140 cm at the 
for bodies such as Venus and Mars compared to previously published 
Apollo 15 and 17 landing sites located at 26oN; 3.6oE and 20oN; 30.6oE, 
data. Our decision to recalculate was based on the assumption that the 
respectively [88]. Due to a strong temperature dependence of the lunar 
orbital distances of planets are known with much greater accuracy than 
regolith thermal conductivity in the topmost 1-2 cm soil, the Moon’s 
TOA solar irradiances  Hence, a correction made to Earth’s irradiance 
average diurnal temperature increases steadily with depth. According 
requires adjusting the ‘solar constants’ of all other planets as wel .
to Apollo measurements, the mean daily temperature at 35 cm 
 OFFICIAL 
belowground is 40–45 K higher than that at the lunar surface [88]. The 
We found that quoted values for the mean global temperature and 
diurnal temperature fluctuations completely vanish below a depth of 80 
surface atmospheric pressure of Mars were either improbable or too 
cm. At 100 cm depth, the temperature of the lunar regolith ranged from 
uncertain to be useful for our analysis. Thus, studies published in the 
250.7 K to 252.5 K at the Apollo 15 site and between 254.5 K and 255.5 K 
last 15 years report Mars’ GMAT being anywhere between 200 K and 
at the Apollo 17 site [88]. Hence, reported Moon average temperatures
240 K with the most frequently quoted values in the range 210–220 
in the range 250-255 K do not describe surface conditions. Moreover,
K [6,32,76-81]. However, in-situ measurements by Viking Lander 1 
since measured in the lunar subtropics, such temperatures do not likely 
suggest that the average surface air temperature at a low-elevation site 
even represent Moon’s global thermal environment at these depths. On
in the Martian subtropics does not exceed 207 K during the summer-
the other hand, frequently quoted Moon global temperatures of ~270 K 
fall season (Appendix B). Therefore, the Red Planet’s GMAT must be 
have actual y been calculated from Eq. (3) and are not based on surface 
lower than 207 K. The Viking records also indicate that average diurnal 
measurements. However, as demonstrated by Volokin and ReLlez [1],
Environ Pollut Climate Change, an open access journal 
Volume 1 • Issue 2 • 1000112

Citation: N kolov N, Zeller K (2017) New Insights on the Physical Nature of the Atmospheric Greenhouse Effect Deduced from an Empirical Planetary 
Temperature Model. Environ Pollut Climate Change 1: 112. 
Page 6 of 22
Eq. (3) overestimates the mean global surface temperature of spheres 
Greenhouse gases in planetary atmospheres represented by the major 
by about 37%. In this study, we employed the spherical estimate of 
constituents carbon dioxide (CO ), methane (CH ) and water vapor 
2
4
Moon’s GMAT (197.35 K) obtained by Volokin and ReLlez [1] using 
(H O) were collectively quantified via three bulk parameters: average 
2
output from a NASA thermo-physical model validated against Diviner 
molar mass (, kg mol-1), combined partial pressure (, Pa) and 
gh
gh
observations [29].
combined partial density (ρ , kg m-3). These parameters were estimated 
gh
Surprisingly, many publications report incorrect values even  from reported volumetric concentrations of individual greenhouse 
for Earth’s mean global temperature. Studies of terrestrial climate 
gases (, %) and data on total atmospheric pressure and density in 
x
typical y focus on temperature anomalies and if Earth’s GMAT is 
Table 2 using the formulas:
ever mentioned, it is often loosely quoted as 15 C (~288 K) [2-4,6]. 
 =
C
+
C
+
C
C
gh
0.044 
0.016 
0.018 
/
 
(
CO2
CH4
H2O )
gh
(7)  
However, observations archived in the HadCRUT4 dataset of the 
P
C
gh
    0.01 
 
(
gh )
UK Met Office’s Hadley Centre [50,89] and in the Global Historical 
(8)  
Climatology Network [51,52,90,91] indicate that, between 1981 and 
ρ = ρ
C
M
M
(9)  
gh
    0.01  gh
gh /
 
(
)(
)
2010, Earth’s mean annual surface air temperature was 287.4 K (14.3 
C) ± 0.5 K. Some recent studies acknowledge this more accurate lower
where  =  is the total volumetric concentration 
gh 
CO2 
CH4 
H2O 
value of Earth’s absolute global temperature [92]. For Earth’s mean
of major greenhouse gases (%). The reference temperatures Τ  and Τ  
e
na
surface atmospheric pressure we adopted the estimate by Trenberth et
were calculated from Equation  (3) and (4c), respectively. 
al. [53] (98.55 kPa), which takes into account the average elevation of
Results
continental landmasses above sea level; hence, it is slightly lower than
the typical sea-level pressure of ≈ 101.3 kPa.
Function (5) was fitted to each one of the 12 sets of logarithmic π   i
pairs gen rated by Equations (1) and (2) and shown in Table 4. Figures 
The average near-surface atmospheric densities (ρ, kg m-3) of 
1 and 2 display the resulting curves of  individual regression models 
ACT 1982
planetary bodies were calculated from reported means of total  with planetary data plotted in the background for reference. Table 5 lists 
atmospheric pressure (P), molar mass (M, kg mol-1) and temperature 
the st tistical scores  f each non-linear regression. Model 12 depicted 
() using the Ideal Gas Law, i.e. 
s
in Figure 2f  had the highest R2 = 0.9999 and the lowest standard error 
P 
  
M
ρ =
(6)
σ  = 0.0078 among all regressions. Model 1 (Figure 1a) provided the 
t
R T
second best fit with R2 = 0.9844 and σ  = 0.1529. Notably, Model 1 
s
est
where  R = 8.31446 J mol-1  K-1  is the universal gas constant. This 
shows almost a 20-time larger standard error on the logarithmic scale 
calculation was intended to make atmospheric densities physical y 
than Model 12. Figure 3 il ustrates the difference in predictive skil s 
consistent with independent data on pressure and temper ture utilized 
between the two top-performing Models 1 and 12 upon conversion 
in our study. The resulting ρ values were similar to previously published 
of vertical axes to a linear scale. Taking an antilogarithm weakens 
data for individual bodies. Standard errors of the air-density estimates 
he relationship of Model 1 to the point of becoming immaterial and 
were calculated from reported errors of P and Τ  fo  each body usin  
highlights the superiority of Model 12. The statistical results shown in 
Eq. (6).
Table 5 indicate that the explanatory power and descriptive accuracy of 
Model 12 surpass those of all other models by a wide margin. 
Data in Table 2 were harnessed to comp te several intermediate 
variables and all dimensionless π  product  necessary for the regression 
Since Titan and Earth nearly overlap on the logarithmic scale of Figure 
i
analyses. The results from these computations are shown in Table 4. 
2f, we decided to experiment with an alternative regression for Model 12, 
INFORMATION 
Intermediate Variable or Dimensionless Product
Venus
Earth
Moon
Mars
Titan
Triton
Average molar mass of greenhouse gases,  (kg mo -1) 
gh
0.0440
0.0216
0.0
0.0440
0.0160
0.0160
(Eq. 7)
RELEASED UNDER THE 
Near-surface partial pressure of greenhouse gases,  (Pa)  8,974,500.0 ± 
gh
(Eq. 8) 
96,500
283.8 ± 0.02
0.0
667.7 ± 13.8
7,188.3 ± 4.9
9.6 × 10-4 ± 2.9 
× 10-4
Near-surface density of greenhouse gases  ρ  (kg m-3) (Eq. 9)
64.441 ± 0.429 2.57 × 10-3 ± 4.3 
0.148 ± 8.4 ×  4.74 × 10-8 ± 1.3 
gh
× 10-6
0.0
0.018 ± 3.1 × 
10-4
10-4
× 10-8
Radiating equilibrium temperature, T  (K) (Eq. 3)
185.0
256.4
269.7
211.9
83.6
39.2
e
Average airless spherical temperature, T  (K) (Eq. 4c)
231.7
197.0
197 0
159.6
63.6
35.9
na 
T / T
3.985 ± 0.016
1.121 ± 0.002
0.732 ± 0.003
0.899 ± 0.003
1.120 ± 0.008
0.994 ± 0 026
s
e
T /T
3.181 ± 0.013
1.459 ± 0.002
1.002 ± 0.004
1.194 ± 0.004
1.473 ± 0.011
1.086 ± 0 028
s
na
ln(/
1.3825 ± 0.0041 0.1141 ± 0.0017 -0.3123 ± 0.0046 -0.1063 ± 0.0037 0.1136 ± 0.0075
-5.2×10-3 ± 
s
e
0.0256
 OFFICIAL 
 ln(/) 
1.1573 ± 0.0041 0.3775 ± 0.0017
1.59×10-3 ± 
s
na
0.0046
0.1772 ± 0.0037 0.3870 ± 0.0075 0.0828 ± 0 0256
ln[3/(ρ  S2)]
28.1364
8.4784
Undefine
10.7520
23.1644
-4.7981
gh
gh
ln[P3/(ρ  S2)]
28 2433
26.0283
+∞
10.8304
32.2122
20.2065
gh
ln[3/(ρ S2)]
28.1145
2.3370
Undefine
10.7396
19.6102
-13.6926
gh
ln[/]
9.5936
-0.7679
Undefine
0.0876
2.4639
-13.3649
gh
r
ln[P3/(ρ S2)]
28 2214
19.8869
-46.7497
10.8180
28.6580
11.3120
ln(P
-28.3570 ± 
-5.0300 ± 
/)
9.6292 ± 0.0108
5.0820 ± 
r
6.6×10-5
0.3516
0.1137 ± 0.0207
5.4799 ± 
6.8×10-4
0.3095
Table 4:  Intermediate  variables  and  dimensionless  products  required  for  the  regression  analyses  and  calculated  from  data  in  Table  2.  Equations  used  to  compute 
intermediate variables are shown in parentheses. The reference pressure is set to the barometric triple point of water, i e. P  = 611.73 Pa.
r
Environ Pollut Climate Change, an open access journal 
Volume 1 • Issue 2 • 1000112







Citation: N kolov N, Zeller K (2017) New Insights on the Physical Nature of the Atmospheric Greenhouse Effect Deduced from an Empirical Planetary 
Temperature Model. Environ Pollut Climate Change 1: 112. 
Page 7 of 22
ACT 1982
INFORMATION 
Figure 1: The  rela ive  atmospheric  thermal  enhancement  (T /T )  as  a  function  of  various  dimensionless  forcing  variables  generated  by  DA  using  data  on  solar 
s
r
irradiance, near-surface partial pressure/den ity of greenhouse gases, and total atmospheric pressure/density from Table 4. Panels a through f depict six regression 
models suggested by DA with the underlying celestial bodies plotted in the background for reference. Each pair of horizontal graphs represents different reference 
RELEASED UNDER THE 
temperatures () defined as either T  = T  ( eft) or T  = T  (right).
r
r
e
r
na  
which excludes Titan from the input dataset. This new curve had R2 = 
Equation (10a) implies that GMATs of rocky planets can be 
1.0 and σ  = 0.0009. Although the two regression equations yield similar 
calculated as a product of two quantities: the planet’s average surface 
est
results over most of the relevant pressure range, we chose the one without 
temperature in the absence of an atmosphere (, K) and a non-
na
Titan as final for Model 12 based on the assumption that Earth’s GMAT 
dimensional factor (≥ 1.0) quantifying the relative thermal effect of 
a 
is likely known with a much greater accuracy than Titan’s mean annual 
the atmosphere, i.e. 
temperature. Taking an antilogarithm of the final regression equation, 
which excludes Titan, yielded the following expression fo
 OFFICIAL r Model 12:
T E
(10b)
s
na   a
0.150263
1.04193


T
 
 
where Τ  is obtained from the SAT model (Eq. 4a) and  is a function 

P
s
5
=  exp 0.174205 
+ 1.83121  ×  10  
               (10a)
na
a
of total pressure (P) given by:
na
T

P
P
 
 

r
r


0.150263
1.04193




The regression coefficients in Eq. (10a) are intentional y shown in 
 



P
E
=





×


   (11)
(
)
5
 exp 0.174205
 exp 1.83121 10
           
full precision to allow an accurate calculation of RATE (i.e. the T /

P




P



r
r
s




  ratios) provided the strong non-linearity of the relationship and 
na
to facilitate a successful replication of our results by other researchers. 
Note that, as P approaches 0 in Eq. (11),  approaches the physical y 
a
Figure 4 depicts Eq. (10a) as a dependence of RATE on the average 
realistic limit of 1.0. Other physical aspects of this equation are 
surface air pressure. Superimposed on this graph are the six planetary 
discussed below.   
bodies from Table 4 along with their uncertainty ranges. 
For bodies with tangible atmospheres (such as Venus, Earth, 
Environ Pollut Climate Change, an open access journal 
Volume 1 • Issue 2 • 1000112







Citation: N kolov N, Zeller K (2017) New Insights on the Physical Nature of the Atmospheric Greenhouse Effect Deduced from an Empirical Planetary 
Temperature Model. Environ Pollut Climate Change 1: 112. 
Page 8 of 22
ACT 1982
 Figure 2: The same as in Figure 1 but for six additional regression models (panels a through f).
INFORMATION 
Mars, Titan and Triton), one must calculate T  using α  = 0.132 and 
surface in a way that transforms these parameters from independent 
n  
e
η  = 0.00971, which assumes a Moon-like airless reference surface in 
controllers of the global temperature in airless bodies to intrinsic 
RELEASED UNDER THE 
e
accordance with our pre-analysis premise  For bodies with tenuous 
byproducts of the climate system itself in worlds with appreciable 
atmospheres (such as Mercury,  he Moon, Calisto and Europa), T   na 
atmospheres. In other words, once atmospheric pressure rises above a 
should be calculated from Eq. (4a) (or Eq. 4b respectively if > 0.15 
certain level, the effects of albedo and ground heat storage on GMAT 
W m-2 and/or Rg ≈ 0 W m-2) using the body’s observed values of Bond 
become implicitly accounted for by Eq. (11). Although this hypothesis 
albedo α  and ground heat storage fraction η . In the context of this 
e
e
requires a further investigation beyond the scope of the present study, 
model, a tangible atmosphere is defined as one that has significantly 
one finds an initial support for it in the observation that, according to 
modified the optical and thermo-physical properties of a planet’s 
data in Table 2, GMATs of bodies with tangible atmospheres do not 
surface compared to an airless environment and/or noticeably  show a physical y meaningful relationship with the amounts of absorbed 
impacted the overall planetary albedo by enabling the formation of 
shortwave radiation determined by albedos. Our discovery for the 
 OFFICIAL 
clouds and haze. A tenuous atmosphere, on the other hand, is one that 
has not had a measurable influence on the surface albedo and regolith 
need to utilize different albedos and heat storage coefficients between 
thermo-physical properties and is completely transparent to shortwave 
airless worlds and worlds with tangible atmospheres is not unique as a 
radiation. The need for such delineation of atmospheric masses when 
methodological approach. In many areas of science and engineering, 
calculating T  arises from the fact that Eq. (10a) accurately describes 
it is sometime necessary to use disparate model parameterizations to 
na 
RATEs of planetary bodies with tangible atmospheres over a wide 
successful y describe different aspects of the same phenomenon. An 
range of conditions without explicitly accounting for the observed large 
example is the distinction made in fluid mechanics between laminar 
differences in albedos (i.e. from 0.235 to 0.90) while assuming constant 
and turbulent flow, where the non-dimensional Reynold’s number is 
values of α  and η  for the airless equivalent of these bodies. One possible 
employed to separate the two regimes that are subjected to different 
e
e
explanation for this counterintuitive empirical result is that atmospheric 
mathematical treatments.
pressure alters the planetary albedo and heat storage properties of the 
Environ Pollut Climate Change, an open access journal 
Volume 1 • Issue 2 • 1000112



Citation: N kolov N, Zeller K (2017) New Insights on the Physical Nature of the Atmospheric Greenhouse Effect Deduced from an Empirical Planetary 
Temperature Model. Environ Pollut Climate Change 1: 112. 
Page 9 of 22
No.
Functional Model
Coefficient of Determination (R2)
Adjusted R2
Standard Error σest
3
T
  P

1
gh
s
 



2
0.9844
0.9375
0.1529
T
 ρ 
 gh  
e
 

3
T
  P

2
gh
s
 



2
0.9562
0.8249
0.1773
T
 ρ 
 gh  
na
 

3
T
 
s
 
3


2
0.1372
-2.4511
1.1360
T
  ρ 
 gh  
e
 

3
T
 
s
 
4


2
0.2450
-2.0200
0.7365
T
  ρ 
 gh  
na
 

3
T
   
5
gh
s
 



2
0.9835
0.9339
0.1572
T
 ρ  
e
S


3
T
   
6
gh
s
 



2
0.9467
0.7866
0.1957
T
 ρ  
na
S


T
 
7
gh
s
 


0.9818
0.927
0.1648
e
T
P
 
T
 
8
gh
s
 
=
ACT 1982


0.9649
0.8598
0.1587
na
T
P
 
3
T
 
s
 
9

2 
0.4488
-0.3780
0.7060
T
  ρ 
e

3
T
 
s
 
10

2 
0.6256
0.0639
0.4049
T
  ρ 
na

T
 
11


0.9396
0.8489
0.2338
e
T
P
 
T
 
12
s
 
0.9999
0.9997
0.0078
na
T
P
 
Table 5: Performance statistics of the twelve regression models suggested by DA. Statistical scores refer to the model logarithmic forms shown in Figures 1 and 2. 
INFORMATION 
RELEASED UNDER THE 
 OFFICIAL 
Figure 3: Comparison of the two best-performing regression models according to statistical scores listed in Table 5. Vertical axes use linear scales to better illustrate 
he difference in skills between the models.
We do not currently have sufficient data to precisely define the limit 
and from Eq. (4a) (or Eq. 4b, respectively) using observed values of α  e
between tangible and tenuous atmospheres in terms of total pressure for 
and η  if P ≤ 10-2 Pa. Equation (4a) should also be employed in cases, 
e
the purpose of this model. However, considering that an atmospheric 
where a significant geothermal flux exists such as on the Galilean moons 
pressure of 1.0 Pa on Pluto causes the formation of layered haze [93], 
of Jupiter due to tidal heating, and/or if S  ≤ 0.15 W m-2. Hence, the 
we surmise that this limit likely lies significantly below 1.0 Pa. In this 
30-year mean global equilibrium surface temperature of rocky planets
study, we use 0.01 Pa as a tentative threshold value. Thus, in the context 
depends in general on five factors: TOA stel ar irradiance (S), a reference
of Eq. (10b), we recommend computing T  from Eq. (4c) if P > 10-2 Pa, 
na 
airless surface albedo (α ), a reference airless ground heat storage fraction
e
Environ Pollut Climate Change, an open access journal 
Volume 1 • Issue 2 • 1000112



Citation: N kolov N, Zeller K (2017) New Insights on the Physical Nature of the Atmospheric Greenhouse Effect Deduced from an Empirical Planetary 
Temperature Model. Environ Pollut Climate Change 1: 112. 
Page 10 of 22
K). Equation (10b) produces 95.18 K for Titan at Saturn’s semi-major 
axis (9.582 AU) corresponding to a solar irradiance S = 14.8 W m-2. This 
estimate is virtual y identical to the 95 K average surface temperature 
reported for that moon by the NASA JPL Voyager Mission website 
[94]. The Voyager spacecraft 1 and 2 reached Saturn and its moons in 
November 1980 and August 1981, respectively, when the gas giant was 
at a distance between 9.52 AU and 9.60 AU from the Sun corresponding 
approximately to Saturn’s semi-major axis [95].
Data acquired by Voyager 1 suggested an average surface 
temperature of 94 ± 0.7 K for Titan, while Voyager 2 indicated a 
temperature close to 95 K [41]. Measurements obtained between 2005 
and 2010 by the Cassini-Huygens mission revealed ≈ 93.4 ± 0.6 K 
s 
[42,43]. Using Saturn’s perihelion (9.023 AU) and aphelion (10.05 AU) 
one can compute Titan’s TOA solar irradiance at the closest and furthest 
approach to the Sun, i.e  16 7 W m-2 and 13.47 W m-2, respectively. 
Inserting these values into Eq. (10b) produces the expected upper and 
Figure 4:  The  relative  atmospheric  thermal  enhancement  (T /T   ratio)  as  a 
lower limit of Titan’s mean global surface temperature according to 
s
na
function of the average surface air pressure according to Eq. (10a) derived from 
our model, i.e. 92.9 K ≤  ≤ 98.1 K. Notably this range encompasses 
s
data representing a broad range of planetary environments in the solar system. 
all current observation-based estimates of Titan’s GMAT. Since both 
Saturn’s  moon Titan  has  been  excluded  from  the  regression  analysis  leading 
to Eq. (10a). Error bars of some bodies are not clearly visible due to their small 
Voyager and C ssini mission covered shorter periods than a single 
size relative to the scale of the axes. See Table 2 for the actual error estimates.
Titan  eason (Saturn’s orbital period is 29.45 Earth years), the available 
ACT 1982
measurements may not well represent that moon’s annual thermal 
cycle. In addition  due to a thermal inertia, Titan’s average surface 
temperature likely lags variations in the TOA solar irradiance caused 
by Saturn’s orbital eccentricity. Thus, the observed 1.45 K discrepancy 
between our independent model prediction and Titan’s current 
best-known GMAT seems to be within the range of plausible global 
temperature fluctuations on that moon. Hence, further observations are 
needed to more precisely constrain Titan’s long-term GMAT.
 Measurements conducted by the Voyager spacecraft in 1989 
indicated a global mean temperature of 38 ± 1.0 K and an average 
atmospheric pressure of 1.4 Pa at the surface of Triton [73].  Even 
though Eq. (10a) is based on slightly different data for Triton (i.e.   = 
s
39 ±1.0 K and P = 4.0 Pa) obtained by more recent stel ar occultation 
measurements [73], employing the Voyager-reported pressure in Eq. 
(10b) produces  = 38.5 K for Triton’s GMAT, a value well within the 
s
uncertainty of the 1989 temperature measurements.
INFORMATION 
The above comparisons indicate that Eq. (10b) rather accurately 
Figure 5: Absolute differences b tween modeled average global temperatures 
describes the observed variation of the mean surface temperature across 
by  Eq.  (10b)  and  observed  GMATs  (Table  2)  for  he  studied  celestial  bodies. 
a wide range of planetary environments in terms of solar irradiance 
RELEASED UNDER THE 
Saturn’s moon Titan represents an independent data point, since it was excluded 
from the regression analysis leading to Eq. (10a).
(from 1.5 W m-2 to 2,602 W m-2), total atmospheric pressure (from 
near vacuum to 9,300 kPa) and greenhouse-gas concentrations (from 
0.0% to over 96% per volume). While true that Eq. (10a) is based on 
(η ), the average geothermal flux reaching the surface (), and the total 
e
g
data from only 6 celestial objects, one should keep in mind that these 
surface atmospheric pressure (P). For planets with tangible atmospheres 
constitute virtual y all bodies in the Solar System meeting our criteria 
(> 10-2 Pa) and a negligible geothermal heating of the surface ( ≈ 0), 
g
for availability and quality of measured data. Although function (5) 
the equilibrium GMAT becomes only a function of two factors: S and 
has 4 free parameters estimated from just 5-6 data points, there are no 
P, i.e. Τ  = 32.44 S0.25(P). The final model (Eq. 10b) can also be cast 
s
α
signs of model overfitting in this case because (a) Eq. (5) represents 
in terms of  as a function of a planet’s distance to the Sun (, AU) by 
a monotonic function of a rigid shape that can only describe well 
 OFFICIAL 
s
au
replacing S in Equations (4a), (4b) or (4c) with 1360.9 -2.
certain exponential pattern as evident from Figures 1 and 2 and the 
au
Environmental scope and numerical accuracy of the new  statistical scores in Table 5; (b) a simple scatter plot of ln (P/P ) vs. ln(T /
r
s
) visibly reveals the presence of an exponential relationship free of 
model
na
data noise; and (c) no polynomial can fit the data points in Figure 2f 
Figure 5 portrays the residuals between modeled and observed 
as accurately as Eq. (5) while also producing a physical y meaningful 
absolute planetary temperatures. For celestial bodies participating in 
response curve similar to known pressure-temperature relationships in 
the regression analysis (i.e. Venus, Earth, Moon, Mars and Triton), the 
other systems. These facts indicate that Eq. (5) is not too complicated 
maximum model error does not exceed 0.17 K and is well within the 
to cause an over-fitting but just right for describing the data at hand. 
uncertainty of observations. The error for Titan, an independent data 
The fact that only one of the investigated twelve non-linear 
point, is 1.45 K or 1.5% of that moon’s current best-known GMAT (93.7 
regressions yielded a tight relationship suggests that Model 12 describes 
Environ Pollut Climate Change, an open access journal 
Volume 1 • Issue 2 • 1000112


Citation: N kolov N, Zeller K (2017) New Insights on the Physical Nature of the Atmospheric Greenhouse Effect Deduced from an Empirical Planetary 
Temperature Model. Environ Pollut Climate Change 1: 112. 
Page 11 of 22
a macro-level thermodynamic property of planetary atmospheres 
heretofore unbeknown to science. A function of such predictive power 
spanning the entire breadth of the Solar System cannot be just a result 
of chance. Indeed, complex natural systems consisting of myriad 
interacting agents have been known to sometime exhibit emergent 
responses at higher levels of hierarchical organization that are amenable 
to accurate modeling using top-down statistical approaches [96]. 
Equation (10a) also displays several other characteristics discussed 
below that lend further support to the above notion. 
Model robustness
Model robustness defines the degree to which a statistical 
relationship would hold when recalculated using a different dataset. To 
test the robustness of Eq. (10a) we performed an alternative regression 
analysis, which excluded Earth and Titan from the input data and 
only utilized logarithmic pairs of T /T  and P/P  for Venus, the Moon, 
s
na
r
Mars and Triton from Table 4. The goal was to evaluate how well the 
Figure 6: Demonstration of the robustness  f Model 12. The solid black curve 
resulting new regression equation would predict the observed mean 
depicts Eq. (10a) based on data from 5 celestial bodies (i.e. Venus, Earth, Moon, 
surface temperatures of Earth and Titan. Since these two bodies occupy 
Mars and Triton). The dashed grey curve portrays Eq. (12a) derived from data of 
a highly non-linear region in Model 12 (Figure 2f), eliminating them 
only 4 bodies (i e. Venus, Moon  Mars and Triton) while excluding Earth and Titan 
from  the  r gre sion  analysis. The  alternative  Eq.  (12b)  predicts  the  observed 
from the regression analysis would leave a key portion of the curve 
GMATs of Ea th and Titan with accuracy greater  han 99% indicating that Model 
ACT 1982
poorly defined. As in all previous cases, function (5) was fitted to the 
12 is statistically robust.
incomplete dataset (omitting Earth and Titan), which yielded the 
following expression:
The above characteristics of Eq. (10a) including dimensional 
0.150275
3.32375


T
 


homogeneity  high predictive accuracy, broad environmental scope of 

P
s
15
= exp 0.174222 
+ 5.25043×10  
                    
(12 )
validity and sta istical robustness indicate that it represents an emergent 
na
T

P
P





r
r


macro-physical model of theoretical significance deserving further 
Substituting the reference temperature T  in Eq. (12a) with its  investigation. This conclusive result is also supported by the physical 
na 
equivalent from Eq. (4c) and solving for  produces 
meaningfulness of the response curve described by Eq. (10a).
s
0.150275
3.32375








Discussion
0.25
P
15

P
(12b)
=
S





×



s
32 44 
exp 0 174222
 exp 5 25043 10
      

P




P


     
r
r




Given the high statistical scores of the new model discussed above, 
It is evident that the regression coefficients in the first exponent term of 
it is important to address its physical significance, potential limitations, 
Eq. (12a) are nearly identical to those in Eq. (10a). This term dominates 
and broad implications for the current climate theory. 
the  T -P relationship over the pressure range 0-400 kPa  ccounting 
s
Similarity of the new model to Poisson’s formula and the SB 
for more than 97.5% of the predicted temperature magnitudes. The 
regression coefficients of the second exponent differ somewhat between 
radiation law
the two formulas causing a diverg nce of calculated RATE values 
The functional response of E
INFORMATION  q. (10a) portrayed in Figure 4 closely 
over the pressure interval 400–9,100 kPa. Th  models converge again 
resembles the shape of the dry adiabatic temperature curve in Figure 
between 9,000 kPa and 9,300 kPa. Figure 6 il ustrates the similarity of 
7a described by the Poisson formula and derived from the First Law of 
RELEASED UNDER THE 
responses between Equations (10a) and (12a) over the pressure range 
Thermodynamics and the Ideal Gas Law [4], i.e.
0–300 kPa with Earth and Titan plotted in the foreground for reference.
R/cp
T
 
=  
          (13)
Equation (12b) reproduces the observed global surface temperature 
T
p
o
 
of Earth with an error of 0 4% (-1.0 K) and that of Titan with an error 
of 1.0% (+0.9 K). For Titan, the error of the new Eq. (12b) is even 
Here,  and  are reference values for temperature and pressure 
o
o
slightly smaller than that of the original model (Eq. 10b). The ability 
typical y measured at the surface, while T and p are corresponding scalars 
of Model 12 to predict Earth’s GMAT with an accuracy of 99.6% using 
in the free atmosphere, and  is the molar heat capacity of air (J mol-1 
p
a relationship inferred from disparate environments such as those 
K-1). For the Earth’s atmosphere, R/c  = 0.286. Equation (13) essential y
p
found on Venus, Moon, Mars and Triton indicates tha
 OFFICIAL t (a) this model  describes the direct effect of pressure p on the gas temperature (T) in
is statistical y robust, and (b) Earth’s temperature is a part of a cosmic 
the absence of any heat exchange with the surrounding environment.
thermodynamic continuum well described by Eq. (10b). The apparent 
Equation (10a) is structural y similar to Eq. (13) in a sense that 
smoothness of this continuum for bodies with tangible atmospheres 
both expressions relate a temperature ratio to a pressure ratio, or more 
(il ustrated in Figure 4) suggests that planetary climates are well-
precisely, a relative thermal enhancement to a ratio of physical forces. 
buffered and have no ‘tipping points’ in reality, i.e. states enabling 
However, while the Poisson formula typical y produces 0 ≤ T/T  ≤ 1.0, 
o
rapid and irreversible changes in the global equilibrium temperature 
Eq. (10a) always yields T /T  ≥ 1.0. The key difference between the two 
s
na
as a result of destabilizing positive feedbacks assumed to operate within 
models stems from the fact that Eq. (13) describes vertical temperature 
climate systems. This robustness test also serves as a cross-validation 
changes in a free and dry atmosphere induced by a gravity-controlled 
suggesting that the new model has a universal nature and it is not a 
pressure gradient, while Eq. (10a) predicts the equilibrium response of a 
product of overfitting.
planet’s global surface air temperature to variations in total atmospheric 
Environ Pollut Climate Change, an open access journal 
Volume 1 • Issue 2 • 1000112



Citation: N kolov N, Zeller K (2017) New Insights on the Physical Nature of the Atmospheric Greenhouse Effect Deduced from an Empirical Planetary 
Temperature Model. Environ Pollut Climate Change 1: 112. 
Page 12 of 22
Figure 7: Known pressure-temperature kinetic relations: (a) Dry adiabatic response of the air/surface temperature ratio to pressure changes in a free dry atmosphere 
according to Poisson’s formula (Eq. 13) with a reference pressure set to p  = 100 kPa; (b) The SB radiation law expressed as a response of a blackbody temperature 
o
ratio to variations in photon pressure (Eq. 14). Note  he qualitative striking similarity of shapes between these curves and the one portrayed in Figure 4 depicting the 
new planetary temperature model (Eq. 10a).
pressure. In essence, Eq. (10b) could be viewed as a predictor of the 
boosts the int rnal kinetic energy and raises its temperature, a process 
reference temperature  in the Poisson formula. Thus, while qualitatively 
known in thermodynamics as compression heating. The direct effect 
o
similar, Equations (10a) and (13) are quantitatively rather different. Both 
of pressure on a system’s temperature is thermodynamical y described 
ACT 1982
functions describe effects of pressure on temperature but in the context of 
by adi batic proce ses  The pressure-induced thermal enhancement 
disparate physical systems. Therefore, estimates obtained from Eq. (10a) 
at a planetary lev l portrayed in Figure 4 and accurately quantified by 
should not be confused with results inferred from the Poisson formula. 
Eq  (10a or 11) is analogous to a compression heating, but not ful y 
For example, Eq. (10b) cannot be expected to predict the temperature 
i entical to an adiabatic process. The latter is usual y characterized by 
lapse rate and/or vertical temperature profiles within a planetary  a limited duration and oftentimes only applies to finite-size parcels of 
atmosphere as could be using Eq. (13). Furthermore, Eq. (10a) represents 
air moving vertical y through the atmosphere. Equation (11), on the 
a top-down empirical model that implicitly accounts for a plethora of 
other hand, describes a surface thermal effect that is global in scope and 
thermodynamic and radiative processes and feedbacks operating in real 
permanent in nature as long as an atmospheric mass is present within 
climate systems, while the Poisson formula (derived from the Ideal Gas 
the planet’s gravitational field. Hence, the planetary RATE (T /T ratio) 
s
na 
Law) only describes pressure-induced temperature changes in a simple 
c uld be understood as a net result of countless simultaneous adiabatic 
mixture of dry gases without any implicit or explicit consideration of 
processes continuously operating in the free atmosphere. Figures 4 and 
planetary-scale mechanisms such as latent heat transport and cloud 
7 also suggest that the pressure control of temperature is a universal 
radiative forcing. 
thermodynamic principle applicable to systems ranging in complexity 
Equation (10a) also shows a remarkable similarity  o the SB law 
from a simple isothermal blackbody absorbing a homogeneous flux of 
relating the equilibrium skin temp rature  f an isotherma  blackbody 
electromagnetic radiation to diverse planetary atmospheres governed 
(, K) to the electromagnetic radiati e flux (I, W m
by complex non-linear process interactions and cloud-radiative 
-2) absorbed/
b
INFORMATION 
emitted by the body’s surface, i .  = (I ⁄ σ)
feedbacks. To our knowledge, this cross-scale similarity among various 
0.25. Dividing each side of 
this fundamental relationship by the irreducible temperature of deep 
pressure-temperature relationships has not previously been identified 
Space   = 2.725 K and its causative CMBR  = 3.13 × 10
and could provide a valuable new perspective on the working of 
-6 W m-2 
c
c
RELEASED UNDER THE 
respectively, yields T ⁄T   = (I ⁄  )
planetary climates.
0.25. Further, expressing the radiative 
b c
c
fluxes I and  on the right-hand side as products of photon pressure 
Nevertheless, important differences exist between Eq. (10a) and 
c
and the speed of light (c, m s-1) in a vacuum, i.e. I = cP  and  R  = cP 
these other  simpler pressure-temperature relations. Thus, while the 
ph
c
c
leads to the following alternative f rm of the SB law:
Poisson formula and the SB radiation law can mathematical y be 
0 25
 
derived from ‘first principles’ and experimental y tested in a laboratory, 
T
 
ph
 
= 

(14)
Eq. (10a) could neither be analytical y deduced from known physical 
T
P
c
 
laws nor accurately simulated in a small-scale experiment. This is 
where  = 1.043 × 10-14 Pa is the photon pressure of CMBR. Clearly, Eq. 
c
because Eq. (10a) describes an emergent  macro-level property of 
(10a) is analogous to Eq. (14), while the latter is structural y identical to 
planetary atmospheres representing the net result of myriad process 
the Poisson formula (13). Figure 7b depicts Eq. (14) as a dep
 OFFICIAL  endence of  interactions within real climate systems that are not readily computable 
the /  ratio on photon pressure 
b
c
ph
using mechanistic (bottom-up) approaches adopted in climate models 
It is evident from Figures 4 and 7 that formulas (10a), (13) and (14) 
or ful y reproducible in a laboratory setting. 
describe qualitatively very similar responses in quantitatively vastly 
Potential limitations of the planetary temperature model
different systems. The presence of such similar relations in otherwise 
disparate physical systems can fundamental y be explained by the fact 
Equation (10b) describes long-term (30-year) equilibrium GMATs 
that pressure as a force per unit area represents a key component of 
of planetary bodies and does not predict inter-annual global temperature 
the internal kinetic energy (defined as a product of gas volume and 
variations caused by intrinsic fluctuations of cloud albedo and/or ocean 
pressure), while temperature is merely a physical manifestation of this 
heat uptake. Thus, the observed 0.82 K rise of Earth’s global temperature 
energy. Adding a force such as gas pressure to a physical system inevitably 
since 1880 is not captured by our model, as this warming was likely 
Environ Pollut Climate Change, an open access journal 
Volume 1 • Issue 2 • 1000112

Citation: N kolov N, Zeller K (2017) New Insights on the Physical Nature of the Atmospheric Greenhouse Effect Deduced from an Empirical Planetary 
Temperature Model. Environ Pollut Climate Change 1: 112. 
Page 13 of 22
not the result of an increased atmospheric pressure. Recent analyses of 
significantly over the past 65.5 My could open exciting new research 
observed dimming and brightening periods worldwide [97-99] suggest 
venues in Earth sciences in general and paleoclimatology in particular.
that the warming over the past 130 years might have been caused by a 
decrease in global cloud cover and a subsequent increased absorption of 
Role of greenhouse gasses from a perspective of the new 
solar radiation by the surface. Similarly, the mega shift of Earth’s climate 
model 
from a ‘hothouse’ to an ‘icehouse’ evident in the sedimentary archives 
 Our analysis revealed a poor relationship between GMAT and the 
over the past 51 My cannot be explained by Eq. (10b) unless caused by 
amount of greenhouse gases in planetary atmospheres across a broad 
a large loss of atmospheric mass and a corresponding significant drop 
range of environments in the Solar System (Figures 1-3 and Table 5). 
in surface air pressure since the early Eocene. Pleistocene fluctuations 
This is a surprising result from the standpoint of the current Greenhouse 
of global temperature in the order of 3.0–8.0 K during the last 2 My 
theory, which assumes that an atmosphere warms the surface of a planet 
revealed by multiple proxies [100] are also not predictable by Eq. (10b) 
(or moon) via trapping of radiant heat by certain gases controlling the 
if due to factors other than changes in total atmospheric pressure and/
atmospheric infrared optical depth [4,9,10]. The atmospheric opacity 
or TOA solar irradiance. 
to LW radiation depends on air density and gas absorptivity, which in 
The current prevailing view mostly based on theoretical  turn are functions of total pressure, temperature  and greenhouse-gas 
considerations and results from climate models is that the Pleistocene 
concentrations [9]. Pressure also controls the broadening of infrared 
glacial-interglacial cycles have been caused by a combination of three 
absorption lines in individual gases. Therefore, the higher the pressure, 
forcing agents: Milankovitch orbital variations, changes in atmospheric 
the larger the infrared optical depth of an atmosphere, and the stronger 
the expected greenhouse effect would be. According to the current 
concentrations of greenhouse gases, and a hypothesized positive ice-
climate theory, pressure only indirectly affects global surface temperature 
albedo feedback [101,102]. However, recent studies have shown that 
through the atmospheric infrared opacity and its presumed constraint on 
orbital forcing and the ice-albedo feedback cannot explain key features 
the planet’s LW emission to Space [9,107].
of the glacial-interglacial oscil ations such as the observed magnitudes 
ACT 1982
of global temperature changes, the skewness of temperature response 
There are four plausible explanations for the apparent lack of a 
(i.e. slow glaciations followed by rapid meltdowns), and the mid-
close relationship between GMAT and atmospheric greenhouse gasses 
Pleistocene transition from a 41 Ky to 100 Ky cycle length [103-105]. The 
in  ur results: 1) The amounts of greenhouse gases considered in our 
only significant forcing remaining in the present paleo-climatological 
analysis only refer to near-surface atmospheric compositions and 
toolbox to explicate the Pleistocene cycles are variations in greenhouse-
do not de cribe the infrared optical depth of the entire atmospheric 
gas concentrations. Hence, it is difficult to explain, from a standpoint 
column; 2) The analysis lumped all greenhouse gases together and did 
of the current climate theory, the high accuracy of Eq. (11) describing 
not take into account differences in the infrared spectral absorptivity of 
the relative thermal effect of diverse planetary atmospheres without any 
individual gasses; 3) The effect of atmospheric pressure on broadening 
consideration of greenhouse gases. If presumed forcing agents such as 
the infrared gas absorption lines might be stronger in reality than 
greenhouse-gas concentrations and the planetary albedo were ind ed 
simulated by current radiative-transfer models, so that total pressure 
responsible for the observed past temperature dynamics on Earth, why 
overrides the effect of a varying atmospheric composition across a wide 
did these agents not show up as predictors of contemporary planetary 
range of planetary environments; and 4) Pressure as a force per unit area 
temperatures in our analysis as well? Could it be because the e agents 
directly impacts the internal kinetic energy and temperature of a system 
have not real y been driving Earth’s climate on geological time scales? 
in accordance with thermodynamic principles inferred from the Gas 
We address the potential role of greenhouse gases in more d tails below. 
Law; hence, air pressure might be the actual physical causative factor 
Since the relationship portrayed in Figure 4 is undoubtedly real, our 
controlling a planet’s surface temperature rather than the atmospheric 
INFORMATION 
model results point toward the need to reexamine some fundamental 
infrared optical depth, which merely correlates with temperature due to 
climate processes thought to be well understood for decades. For 
its co-dependence on pressure.
example, we are currently testing a hypo hesi  that Pleistocene glacial 
Based on evidence discussed earlier, we argue that option #4 is 
RELEASED UNDER THE 
cycles might have been caused by variations in Earth’s total atmospheric 
the most likely reason for the poor predictive skill of greenhouse 
mass and surface air pressure. Preliminary results based on the ability 
gases with respect to planetary GMATs revealed in our study (Figures 
of an extended version of our planetary model (simulating meridional 
1-3). By definition, the infrared optical depth of an atmosphere is a
temperature gradients) to predict the observed polar amplification 
dimensionless quantity that carries no units of force or energy [3,4,9].
during the Last Glacial Maximum indicate that such a hypothesis is not 
Therefore, it is difficult to fathom from a fundamental physics standpoint 
unreasonable. However  conclusive findings from this research will be 
of view, how this non-dimensional parameter could increase the kinetic 
discussed elsewhere.
energy (and temperature) of the lower troposphere in the presence of
free convection provided that the latter dominates the heat transport in 
According to the present understanding, Earth’s atmospheric  gaseous systems. Pressure, on the other hand, has a dimension of force
pressure has remained nearly invariant during the Cen
 OFFICIAL  ozoic era (last  per unit area and as such is intimately related to the internal kinetic
65.5 My). However, this notion is primarily based on theoretical 
energy of an atmosphere E (J) defined as the product of gas pressure (P,
analyses [106], since there are currently no known geo-chemical proxies 
Pa) and gas volume (V, m3), i.e. E (J) = PV. Hence, the direct effect of
permitting a reliable reconstruction of past pressure changes in a 
pressure on a system’s internal energy and temperature follows straight
manner similar to that provided by various temperature proxies such as 
from fundamental parameter definitions in classical thermodynamics.
isotopic oxygen 18, alkenones and TEX  in sediments, and Ar-N isotope 
86
General y speaking, kinetic energy cannot exist without a pressure
ratios and deuterium concentrations in ice. The lack of independent 
force. Even electromagnetic radiation has pressure.
pressure proxies makes the assumption of a constant atmospheric mass 
throughout the Cenozoic a priori and thus questionable. Although 
In climate models, the effect of infrared optical depth on surface 
this topic is beyond the scope of our present study, allowing for the 
temperature is simulated by mathematical y decoupling radiative 
possibility that atmospheric pressure on Earth might have varied 
transfer from convective heat exchange. Specifical y, the LW 
Environ Pollut Climate Change, an open access journal 
Volume 1 • Issue 2 • 1000112

Citation: N kolov N, Zeller K (2017) New Insights on the Physical Nature of the Atmospheric Greenhouse Effect Deduced from an Empirical Planetary 
Temperature Model. Environ Pollut Climate Change 1: 112. 
Page 14 of 22
radiative transfer is calculated in these models without simultaneous 
is fundamental y different from the hypothesized ‘trapping’ of LW 
consideration of sensible- and latent heat fluxes in the solution matrix. 
radiation by atmospheric trace gases first proposed in the 19th century 
Radiative transfer modules compute the so-called heating rates (K/
and presently forming the core of the Greenhouse climate theory. 
day) strictly as a function of atmospheric infrared opacity, which 
However, a radiant-heat trapping by freely convective gases has never 
under constant-pressure conditions solely depends on greenhouse-
been demonstrated experimental y. We should point out that the hereto 
gas concentrations. These heating rates are subsequently added to the 
deduced adiabatic (pressure-controlled) nature of the atmospheric 
thermodynamic portion of climate models and distributed throughout 
thermal effect rests on an objective analysis of vetted planetary 
the atmosphere. In this manner, the surface warming becomes a 
observations from across the Solar System and is backed by proven 
function of an increasing atmospheric infrared opacity. This approach to 
thermodynamic principles, while the ‘trapping’ of LW radiation by an 
modeling of radiative-convective energy transport rests on the principle 
unconstrained atmosphere surmised by Fourier, Tyndall and Arrhenius 
of superposition, which is only applicable to linear systems, where the 
in the 1800s was based on a theoretical conjecture. The la ter has later 
overall solution can be obtained as a sum of the solutions to individual 
been coded into algorithms that describe the surface temperature as a 
system components. However, the integral heat transport within a 
function of atmospheric infrared optical depth (instead of pressure) by 
free atmosphere is inherently nonlinear with respect to temperature. 
artificial y decoupling radiative transfer from convective heat exchange. 
This is because, in the energy balance equation, radiant heat transfer 
Note also that the Ideal Gas Law (PV = nRT) forming the basis of 
is contingent upon power gradients of absolute temperatures, while 
atmospheric physics is indifferent to the gas chemical composition. 
convective cooling/heating depends on linear temperature differences 
Effect of pressure on temperature: Atmospheric pressure 
in the case of sensible heat flux and on simple vapor pressure gradients 
provides in and of i self only a  elative thermal enhancement (RATE) 
in the case of latent heat flux [4]. The latent heat transport is in turn 
to the surface quantified by Eq. (11). The absolute thermal effect of an 
a function of a solvent’s saturation vapor pressure, which increases 
atmosphere depends on both pressure and the TOA solar irradiance. 
exponential y with temperature [3]. Thus, the superposition principle 
For example, at a total air pressure of 98.55 kPa, Earth’s RATE is 1.459, 
ACT 1982
cannot be employed in energy budget calculations. The artificial  which keeps our planet 90 4 K warmer in its present orbit than it would 
decoupling between radiative and convective heat-transfer processes 
be in the absence of  n atmosphere. Hence, our model ful y explains 
adopted in climate models leads to mathematical y and physical y 
the new ~90 K estimate of Earth’s atmospheric thermal effect derived 
incorrect solutions with respect to surface temperature. The LW  by Volokin and ReLlez [1] using a different line of reasoning. If one 
radiative transfer in a real climate system is intimately intertwined 
moves Earth to the orbit of Titan (located at ~9.6 AU from the Sun) 
with turbulent convection/advection as both transport mechanisms 
without changing the overall pressure, our planet’s RATE will remain 
occur simultaneously. Since convection (and especial y the moist one) 
the same, but the absolute thermal effect of the atmosphere would drop 
is orders of magnitude more efficient in transferring energy than LW 
to about 29.2 K due to a vastly reduced solar flux. In other words, the 
radiation [3,4], and because heat preferential y travel  along the path 
absolute effect of pressure on a system’s temperature depends on the 
of least resistance, a properly coupled radiative-conv ctive algorithm 
background energy level of the environment. This implies that the 
of energy exchange will produce quantitatively and qualitatively  absolute temperature of a gas may not follow variations of pressure 
different temperature solutions in response to a changing atmospheric 
if the gas energy absorption changes in opposite direction to that of 
composition than the ones obtained by current climate models.  pressure. For instance, the temperature of Earth’s stratosphere increases 
Specifical y, a correctly coupled convective-radiative system will render 
with altitude above the tropopause despite a falling air pressure, because 
the surface temperature insensitive to v riations in the atmospheric 
the absorption of UV radiation by ozone steeply increases with height, 
infrared optical depth, a result indir ctly supported by our analysis as 
thus offsetting the effect of a dropping pressure. If the UV absorption 
wel . This topic requires furth r investigation beyond the scope of the 
were constant throughout the stratosphere, the air temperature would 
INFORMATION 
present study. 
decrease with altitude. 
The direct effect of atmospheric pressure on the global surface 
Atmospheric back radiation and surface temperature: Since 
RELEASED UNDER THE 
temperature has received virtual y no a tention in climate science thus 
(according to Eq. 10b) the equilibrium GMAT of a planet is mainly 
far. However, the results from our empirical data analysis suggest that it 
determined by the TOA solar irradiance and surface atmospheric 
deserves a serious consideration in the future. 
pressure, the down-welling LW radiation appears to be global y a product 
of the air temperature rather than a driver of the surface warming. In 
Theoretical implications of the new interplanetary  other words, on a planetary scale, the so-called back radiation is a 
relationship
consequence of the atmospheric thermal effect rather than a cause for 
The hereto discovered pressure-temperature relationship quantified 
it. This explains the broad variation in the size of the observed down-
by Eq. (10a) and depicted in Figure 4 has broad theoretical implications 
welling LW flux among celestial bodies irrespective of the amount of 
that can be summarized as follows: 
absorbed solar radiation. Therefore, a change in this thermal flux brought 
 OFFICIAL 
about by a shift in atmospheric LW emissivity cannot be expected to 
Physical nature of the atmospheric ‘greenhouse effect’: According 
impact the global surface temperature. Any variation in the global 
to Eq. (10b), the heating mechanism of planetary atmospheres is 
infrared back radiation caused by a change in atmospheric composition 
analogous to a gravity-controlled adiabatic compression acting upon 
would be compensated for by a corresponding shift in the intensity of 
the entire surface. This means that the atmosphere does not function 
the vertical convective heat transport. Such a balance between changes 
as an insulator reducing the rate of planet’s infrared cooling to space as 
in atmospheric infrared heating and the upward convective cooling at 
presently assumed [9,10], but instead adiabatical y boosts the kinetic 
the surface is required by the First Law of Thermodynamics. However, 
energy of the lower troposphere beyond the level of solar input through 
current climate models do not simulate this compensatory effect of 
gas compression. Hence, the physical nature of the atmospheric  sensible and latent heat fluxes due to an improper decoupling between 
‘greenhouse effect’ is a pressure-induced thermal enhancement  radiative transfer and turbulent convection in the computation of total 
(PTE) independent of atmospheric composition. This mechanism  energy exchange.
Environ Pollut Climate Change, an open access journal 
Volume 1 • Issue 2 • 1000112

Citation: N kolov N, Zeller K (2017) New Insights on the Physical Nature of the Atmospheric Greenhouse Effect Deduced from an Empirical Planetary 
Temperature Model. Environ Pollut Climate Change 1: 112. 
Page 15 of 22
Role of planetary albedos: The fact that Eq. (10b) accurately 
atmospheric pressure and an extremely cold environment found on that 
describes planetary GMATs without explicitly accounting for the  moon. Thus, our analysis did not reveal evidence for the existence of a 
observed broad range of albedos, i.e. from 0.136 to 0.9 (Table 2), 
feedback between planetary GMAT and a precipitable liquid solvent on 
indicates that the shortwave reflectivity of planetary atmospheres is 
the surface as predicted by the current climate theory. Consequently, 
mostly an intrinsic property (a byproduct) of the climate system itself 
the hypothesized runaway greenhouse, which requires a net positive 
rather than an independent driver of climate as currently believed. In 
feedback between global surface temperature and the atmospheric LW 
other words, it is the internal energy of the atmosphere maintained by 
opacity controlled by water vapor [117], appears to be a model artifact 
solar irradiance and air pressure that controls the bulk of the albedo. 
rather than an actual physical possibility. Indeed, as il ustrated in Figure 
An indirect support for this unorthodox conclusion is provided by 
4, the hot temperature of Venus often cited as a product of a ‘runaway 
the observation that the amounts of absorbed shortwave radiation 
greenhouse’ scenario [117,118] fits perfectly within the pressure-
determined by albedos show no physical y meaningful relationship 
dependent climate continuum described by Equations (10 ) and (11).
with planetary GMATs. For example, data in Table 2 indicate that 
Venus absorbs 3.7 times less solar energy per unit area than Earth, yet 
Model Application and Validation 
its surface is about 450 K hotter than that of Earth; the Moon receives 
Encouraged by the high predictive skill and broad scope of validity 
on average 54 W m-2 more net solar radiation than Earth, but it is 
of Model 12 (Figure 2f) we decided to apply Eq. (10b) to four celestial 
about 90 K cooler on average than our planet. The hereto proposed 
bodies spanning the bread h of the Solar System, i.e. Mercury, Europa, 
passive nature of planetary albedos does not imply that the global 
Callisto and Pluto  which global surface temperatures are not currently 
cloud cover could not be influenced by an external forcing such as solar 
known with certainty. Each body is the target of either ongoing or 
wind, galactic cosmic rays, and/or gravitational fields of other celestial 
planned robotic exploration missions scheduled to provide surface 
objects. Empirical evidence strongly suggests that it can [108-113], but 
thermal d ta among other observations, thus offering an opportunity 
the magnitude of such influences is expected to be small compared to 
to validate our planetary temperature model against independent 
ACT 1982
the total albedo due to the presence of stabilizing negative feedbacks 
measurements. 
within the system. We also anticipate that the sensitivity of GMATs to 
an albedo change will greatly vary among planetary bodies. Viewing 
The MESSENGER spacecraft launched in 2004 completed the first 
the atmospheric reflectivity as a byproduct of the available internal 
comprehensive mapping of Mercury in March 2013 (http://messenger.
energy rather than a driver of climate can also help explain the observed 
jhuapl.edu/).  Among  other  things,  the  spacecraft  also  took  infrared 
remarkable stability of Earth’s albedo [54,114]. 
measurements  of  the  planet’s  surface  using  a  special  spectrometer 
Climate stability: Our semi-empirical model (Equations 4a, 10b 
[119] th t should soon become available. The New Horizons spacecraft
and 11) suggests that, as long as the mean annual TOA solar flux and 
launched  in  January  2006 [120]  reached  Pluto  in  July  of  2015  and
the total atmospheric mass of a planet are stationary, the eq ilibrium 
performed  a  thermal  scan  of  the  dwarf  planet  during  a  flyby.  The
GMAT will remain stable. Inter-annual and decadal variations of global 
complete dataset from this flyby were received on Earth in October of
temperature forced by fluctuations of cloud cove , for exampl , ar  
2016 and are currently being analyzed. A proposed joint Europa-Jupiter
expected to be small compared to the magnitude of the background 
System Mission by NASA and the European Space Agency is planned to
atmospheric warming because of strong neg tive feedbacks limiting 
study the Jovian moons after year 2020. It envisions exploring Europa’s
the albedo changes. This implies a rela ively stable clim te for a planet 
physical and thermal environments both remotely via a NASA Orbiter
such as Earth absent significant shifts in the total  tmospheric mass 
and in situ by a Europa Lander [121].
and the planet’s orbital distance to the Sun. Hence, planetary climates 
All four celestial bodies have somewhat eccentric orbits around the 
appear to be free of tipping points, i.e. functional states fostering 
INFORMATION 
Sun.  However,  while  Mercury’s  orbital  period  is  only  88  Earth  days, 
rapid and irreversible change  in the global temperature as a result of 
Europa  and  Callisto  circumnavigate  the  Sun  once  every  11.9  Earth 
hypothesized positive feedbacks thought to operate within the system. 
years while Pluto takes 248 Earth years. The atmospheric pressure on 
RELEASED UNDER THE 
In other words, our results suggest that the Earth’s climate is well 
Pluto is believed to vary between 1.0 and 4.0 Pa over the course of its 
buffered against sudden changes.
orbital period as a function of insolation-driven sublimation of nitrogen 
Effect of oceans and water vapor on global temperature: The new 
and methane ices on the surface [122]. Each body’s temperature was 
model shows that the Earth’s global equilibrium temperature is a part 
evaluated at three orbital distances from the Sun: aphelion, perihelion, 
of a cosmic thermodynamic continuum controlled by atmospheric 
and the semi-major axis. Since Mercury, Europa and Callisto harbor 
pressure and total solar irradiance. Since our planet is the only one 
tenuous atmospheres (<< 10-2 Pa), the reference temperature T  in 
na 
among studied celesti l bodies harboring a large quantity of liquid 
Eq. (10b) must be calculated from Eq. (4a), which requires knowledge 
water on the surface, Eq. (10b) implies that the oceans play virtual y no 
of the actual values of α η , and . We assumed that Mercury had  = 
role in determining Earth’s GMAT. This finding may sound inexplicable 
e
e
g
g
0.0 W m-2, α  = 0.068 [123] and Moon-like thermo-physical properties 
 OFFICIAL 
from the standpoint of the radiative Greenhouse theory, but it follows 
e
of the regolith (η  = 0.00971). Input data for Europa and Callisto were 
logical y from the new paradigm of a pressure-induced atmospheric 
e
obtained from Spencer et al. [124] and Moore et al. [125], respectively. 
warming. The presence of liquid water on the surface of a planet requires 
Specifical y, in order to calculate η  and  for these moons we utilized 
an air pressure greater than 612 Pa and an ambient temperature above 
e
g
equatorial temperature data provided by Spencer et al. [124] in their 
273.2 K. These conditions are provided by the planet’s size and gravity, 
Figure  1,  and  by  Moore  et  al. [125]  in  their Fig. 17.7  along  with  a 
its distance to the Sun, and the mass of the atmosphere. Hence, the 
theoretical  formula  for  computing  the  average  nighttime  surface 
water oceans on Earth seem to be a thermodynamic consequence of 
temperature T at the equator based on the SB law, i.e.
particular physical conditions set by cosmic arrangements rather than 
an active controller of the global climate. Similarly, the hydrocarbon 
 ( −α )
0.25
 1
 η + 
e
g
=
lakes on the surface of Titan [115,116] are the result of a high 
 

0.98 σ

(15)


Environ Pollut Climate Change, an open access journal 
Volume 1 • Issue 2 • 1000112

Citation: N kolov N, Zeller K (2017) New Insights on the Physical Nature of the Atmospheric Greenhouse Effect Deduced from an Empirical Planetary 
Temperature Model. Environ Pollut Climate Change 1: 112. 
Page 16 of 22
where S(1-α)η  is the absorbed solar flux (W m-2) stored as heat into 
wel  be within the uncertainty of Pluto’s true global temperature. We 
e
the subsurface. The geothermal heat flux on Europa is poorly known. 
will  know  more  about  this  in  2017  when  spatially  resolved  thermal 
However, based on thermal observations of Io reported by Veeder et al. 
measurements obtained by New Horizons become available. 
[126], we assumed  = 2.0 W m-2 for Europa. Using S = 50.3 W m-2, an 
g
observed nighttime equatorial temperature T = 90.9 K and an observed 
One  should  use  caution  when  comparing  results  from  Eq. (10b) 
average night-side albedo α = 0.58 [124], we solved Eq. (15) for the 
to  remotely  sensed ‘average  temperatures’  commonly  quoted  for 
surface heat storage fraction to obtain η  = 0.085 for Europa. A similar 
celestial bodies with tenuous atmospheres such as the moons of Jupiter 
e
computational procedure was employed for Callisto using α = 0.11 and 
and  Neptune.  Studies  oftentimes  report  the  so-called ‘brightness 
equatorial surface temperature data from Fig. 17.7 in Moore et al. [125]. 
temperatures’  retrieved  at  specific  wavelengths  that  have  not  been 
This produced  = 0.5 W m-2 and η  = 0.057. Using these values in 
subjected to a proper spherical integration. As pointed out by Volokin 
g
e
Eq. (15)  correctly  reproduced  Callisto’s  nighttime  equatorial  surface 
and ReLlez [1], due to Hölder’s inequality between integrals, calculated 
temperature of ≈ 86.0 K. The much higher η  estimates for Europa and 
brightness temperatures of spherical objects can be significantly higher 
e
Callisto compared to η  = 0.00971 for the Moon can be explained with 
than actual mean kinetic temperatures of the surface  Sin e Eq. (10b) 
e
the  large  water-ice  content  on  the  surface  of  these  Galilean  moons. 
yields  spherically  averaged  temperatures,  its  predictions  for  airless 
Europa is almost completely covered by a thick layer of water ice, which 
bodies  are  expected  to  be  lower  than  the  disk integrated  brightness 
has a much higher thermal conductivity than the dry regolith. Also, 
temperatures typical y quoted in the literature.
sunlight penetrates deeper into ice than it does into powdered regolith. 
Conclusion
All this enables a much larger fraction of the absorbed solar radiation to 
be stored into the subsurface as heat and later released at night boosting 
For 190 years the atmosphere has been thought to warm Earth 
the nighttime surface temperatures of these moons. Volokin and ReLlez 
by absorbing a portion of the outgoing LW infrared radiation and 
[1] showed that GMAT of airless bodies is highly sensitive to η .
reemitting it back toward the surface, thus augmenting the incident 
e
solar flux  This conceptualized continuous absorption and do
ACT 1982wnward 
Table  6  lists  the  average  global  surface  temperatures  of  the  four
reemission of thermal radiation enabled by certain trace gases known 
celestial bodies predicted by Eq. (10b) along with the employed input 
to be transparent to solar rays while opaque to electromagnetic 
data. According to our model, Mercury is about 117 K cooler on average 
long wavelengths ha  been likened to the trapping of heat by glass 
than NASA’s current estimate of 440 K [32], which is based on Eq. (3) 
greenhouses, hence the term ‘atmospheric greenhouse effect’. Of course, 
and does not represent a spherically averaged surface temperature [1]  
we now know that real greenhouses preserve warmth not by trapping 
Our prediction of Europa’s GMAT, 99.4 K, agrees wel  with the ≈ 100 
infrared radiation but by physical y obstructing the convective heat 
K estimate reported for this moon by Sotin et al. [127]. Our estimate 
exchange between a greenhouse interior and the exterior environment. 
of Pluto’s average surface temperature at perihelion (38.6 K) is similar 
Nevertheless, the term ‘greenhouse effect’ stuck in science. 
to the mean temperature computed for that dwarf planet by Olkin et 
al. [124]  using  a  mechanistic  model  of  nitrogen  ice  volatilization  at 
The hypothesis that a freely convective atmosphere could retain 
the surface. Stern et al. [128] and Gladstone et al. [93] reported initi l 
(trap) radiant heat due its opacity has remained undisputed since its 
results from flyby observations of Pluto taken by the Radio Experiment 
introduction in the early 1800s even though it was based on a theoretical 
(REX) instrument aboard the New Horizons spacecraft in July 2015, 
conjecture that has never been proven experimental y. It is important to 
when  the  dwarf  planet  was  approximately  at  32.9  AU  from  the  Sun. 
note in this regard that the well-documented enhanced absorption of 
Using the observed surface pressure of 1 05 ± 0.1 Pa (10 5 ± 1 μbar) 
thermal radiation by certain gases does not imply an ability of such gases 
[93] our model predicts an average global temperature of 36.7 K for
to trap heat in an open atmospheric environment. This is because, in 
Pluto. Stern et al. [128] repor ed a near-surface temperature of ≈ 38
gaseous systems, heat is primarily transferred (dissipated) by convection 
INFORMATION 
K. However, this value was calculat d from pre-flyby global brightness
(i.e. through fluid motion) rather than radiative exchange. If gases of 
measurements rather than deriv d via spherical integration of spatially
high LW absorptivity/emissivity such as CO , methane and water vapor 
2
resolved surface temperatures (Stern, personal communication). Since
were indeed capable of trapping radiant heat, they could be used as 
RELEASED UNDER THE 
global  brightness  temperatures  tend  t   be  higher  than  spherically
insulators. However, practical experience has taught us that thermal 
averaged kinetic surface temperatures [1], our model prediction may
radiation losses can only be reduced by using materials of very low LW 
α  (fraction) 
Predicted Average Global 
Surface Atmospheric 
e
Surface Temperature at Specific Orbital Distances from the Sun
Pressure (Pa)
η  (fraction) 
e 
R  (W m 2)
Aphelion
Semi-major Axis
Perihelion
g
α  = 0.068
e
296.8 K 
323.3 K 
359.5 K 
Mercury
5 × 10-10
η  = 0.00971
R  = 0.0
(0.459 AU)
(0.387 AU)
(0.313 AU)
g
 OFFICIAL  α = 0.62
e
98.1 K 
99.4 K 
100.7 K 
Europa
10-7
η  = 0.085
e
R  = 2.0
(5.455 AU)
(5.203 AU)
(4.951 AU)
g
α  = 0.11
e
101.2 K 
103.2 K 
105.4 K 
Callisto
7.5 × 10-7
η  = 0.057
e
R  = 0.5
(5.455 AU)
(5.203 AU)
(4.951 AU)
g
α  = 0.132
e
30.0 K  
33.5 K  
38.6 K 
Pluto
1.05
η  = 0.00971
R  = 0.0
(49.310 AU)
(39.482 AU)
(29.667 AU)
g
Table 6: Average global surface temperatures predicted by Eq. (10b) for Mercury, Europa, Calisto and Pluto. Input data on orbital distances (AU) and total atmospheric 
pressure (Pa) were obtained from the NASA Solar System Exploration [48] website, the NASA Planetary Factsheet [32] and Gladstone et al. [93]. Solar irradiances required 
by Eq. (10b) were calculated from reported orbital distances as explained in the text. Values of α , η  and  for Europa and Callisto were estimated from observed data by 
e
e
g
Spencer et al. [124] and Moore et al. [125] respectively (see text for details). 
Environ Pollut Climate Change, an open access journal 
Volume 1 • Issue 2 • 1000112

Citation: N kolov N, Zeller K (2017) New Insights on the Physical Nature of the Atmospheric Greenhouse Effect Deduced from an Empirical Planetary 
Temperature Model. Environ Pollut Climate Change 1: 112. 
Page 17 of 22
absorptivity/emissivity and correspondingly high thermal reflectivity 

The ‘greenhouse effect’ is not a radiative phenomenon driven
such as aluminum foil. These materials are known among engineers at 
by the atmospheric infrared optical depth as presently believed, 
NASA and in the construction industry as radiant barriers [129]. It is 
but a pressure-induced thermal enhancement analogous to
also known that high-emissivity materials promote radiative cooling. 
adiabatic heating and independent of atmospheric composition;
Yet, all climate models proposed since 1800s are built on the premise 
that the atmosphere warms Earth by limiting radiant heat losses of the 

The down-welling LW radiation is not a global driver of surface 
surface through the action of infrared absorbing gases aloft.
warming as hypothesized for over 100 years but a product of
the near-surface air temperature controlled by solar heating
If a trapping of radiant heat occurred in Earth’s atmosphere, the 
and atmospheric pressure;
same mechanism should also be expected to operate in the atmospheres 
of other planetary bodies. Thus, the Greenhouse concept should be able 

The albedo of planetary bodies with tangible atmospheres is not 
to mathematical y describe the observed variation of average planetary 
an independent driver of climate but an intrinsic property (a
surface temperatures across the Solar System as a continuous function 
byproduct) of the climate system itself. This does not mean that 
of the atmospheric infrared optical depth and solar insolation. However, 
the cloud albedo cannot be influenced by external forcing such
to our knowledge, such a continuous description (model) does not 
as solar wind or galactic cosmic rays. However, the magnitude
exist. Furthermore, measured magnitudes of the global down-welling 
of such influences  s expected to be small due to the stabilizing
LW flux on planets with thick atmospheres such as Earth and Venus 
effect of negative feedbacks operating within the system. This
indicate that the lower troposphere of these bodies contains internal 
understanding explains the observed remarkable stability of
kinetic energy far exceeding the solar input [6,12,14]. This fact cannot 
planetary albedos;
be explained via re-radiation of absorbed outgoing thermal emissions 

The equilibrium surface temperature of a planet is bound to
by gases known to supply no additional energy to the system. The desire 
remain stable (i.e. within ± 1 K) as long as the atmospheric
to explicate the sizable energy surplus evident in the tropospheres of 
mass and the TOA mean solar irradiance are stationary. Hence, 
ACT 1982
some terrestrial planets provided the main impetus for this research.
Earth’s climate system is well buffered against sudden changes
We combined high-quality planetary data from the last three 
and has n  tipping points;
decades with the classical method of dimensional analysis to search for 
The proposed net positive feedback between surface
an empirical model that might accurately and meaningful y describe 
temperature and the atmospheric infrared opacity controlled
the observed variation of global surface temperatures throughout the 
by water vapor appears to be a model artifact resulting from
Solar System while also providing a new perspective on the nature of  he 
a mathematical decoupling of the radiative-convective heat
atmospheric thermal effect. Our analysis revealed that the equilibrium 
transfer rather than a physical reality.
global surface temperatures of rocky planets with tangible a mospheres 
and a negligible geothermal surface heating can reli bly be estimated 
The hereto reported findings point toward the need for a paradigm 
across a wide range of atmospheric compositions and radiative regimes 
shift in our understanding of key macro-scale atmospheric properties and 
using only two forcing variables: TOA solar irradiance and total surf ce 
processes. The implications of the discovered planetary thermodynamic 
atmospheric pressure (Eq. 10b with T  computed from Eq. 4c). 
relationship (Figure 4, Eq. 10a) are fundamental in nature and require 
na 
Furthermore, the relative atmospheric thermal enhancement (RATE) 
careful consideration by future research. We ask the scientific community 
defined as a ratio of the planet’s actual global surface temperature to 
to keep an open mind and to view the results presented herein as a possible 
the temperature it would have had in the ab ence of a mosphere is ful y 
foundation of a new theoretical framework for future exploration of 
explicable by the surface air pressure  lone (Eq. 10a and Figure 4). At 
climates on Earth and other worlds. 
the same time, greenhouse-gas  oncentrations and/or partial pressures 
INFORMATION 
Appendices
did not show any meaningful relationship to surface temperatures 
across a broad span of plan tary environments considered in our study 
Appendix A. Construction of the Dimensionless π Variables
RELEASED UNDER THE 
(see Figures 1 and 2 and Table 5). 
Table 1 lists 6 generic variables (T , T , S, P , P  and ρ ) composed of 
s
r
x
r
x
Based on statistical criter a including numerical accuracy,  4 fundamental dimensions: mass [M], length [L], time [T], and absolute 
robustness, dimensional homogeneity and a broad environmental  temperature [Θ]. According to the Buckingham Pi theorem [27], this 
scope of validity, the new relationship (Figure 4) quantified by Eq. (10a) 
implies the existence of two dimensionless π  products per set. To 
i
appears to describe an emergent macro-level thermodynamic property 
derive the π  variables we employed the following objective approach. 
i
of planetary atmospheres heretofore unbeknown to science. The  First, we hypothesized that a planet’s GMAT () is a function of all 5 
s
physical significance of this empirical model is further supported by its 
independent variables listed in Table 1, i.e.
striking qualitative resemblance to the dry adiabatic temperature curve 
= ƒ T , S , P , P , ρ
            (A.1)
s
( r
x
r
)
described by the Poisson formula (Eq. 13) and to the photon-pressure 
 OFFICIAL 
form of the SB radiation law (Eq. 14). Similar to these well-known 
This unknown function is described to a first approximation as a simple 
kinetic relations, Eq. (10a) also predicts the direct effect of pressure on 
product of the driving variables raised to various powers, i.e.
temperature albeit in the context of a different macro-physical system. 
a
b
c
d
e
≈ T S P P ρ
            (A.2)
To our knowledge, this is the first model accurately describing the 
s
r
x
r
x
average surface temperatures of planetary bodies throughout the Solar 
where abcd and e are rational numbers. In order to determine the 
System in the context of a thermodynamic continuum using a common 
power coefficients, Eq. (A.2) is cast in terms of physical dimensions of 
set of drivers.
the participating variables, i.e.

b


c


d

e
a
The planetary temperature model consisting of Equations (4a), 
[Θ] ≈ [Θ]
3
1
2
1
2
3
 M T   M L  T   M L  T   M L 

 
 
 
      (A.3)
(10b), and (11) has several fundamental theoretical implications, i.e. 
Satisfying the requirement for dimensional homogeneity of Eq. 
Environ Pollut Climate Change, an open access journal 
Volume 1 • Issue 2 • 1000112



Citation: N kolov N, Zeller K (2017) New Insights on the Physical Nature of the Atmospheric Greenhouse Effect Deduced from an Empirical Planetary 
Temperature Model. Environ Pollut Climate Change 1: 112. 
Page 18 of 22
(A.2) implies that the sum of powers of each fundamental dimension 
Appendix B. Estimation of Mars’ GMAT and Surface 
must be equal on both sides of Eq. (A.3). This allows us to write four 
Atmospheric Pressure
simultaneous equations (one per fundamental dimension) containing 
five unknowns, i.e.
Although Mars is the third most studied planetary body in the 
Solar System after Earth and the Moon, there is currently no consensus 
 = 1
: [Θ]     
among researchers regarding its mean global surface temperature (). 

M
 = 0
: []     
 values reported over the past 15 years span a range of 40 K. Examples 

M
− − 3= 0
: [L]     

of disparate GMATs quoted for the Red Planet include 200 K [79], 202 
  3
− − 2− 2=

0             : []       
               (A.4)
K [82,130], 210 K [32], 214 K [80], 215 K [6,81], 218 K [77], 220 K [76], 
227 K [131] and 240 K [78]. The most frequently cited temperatures fall 
System (A.4) is underdetermined and has the following solution: 
between 210 K and 220 K. However, a close examination of the available 
= 1, b = 2e, and c  =  -(3e + d). Note that, in the DA methodology, 
thermal observations reveals a high improbability for any of the above 
one oftentimes arrives at underdetermined systems of equations,  estimates to represent Mars’ true GMAT.
simply because the number of independent variables usual y exceeds 
the number of fundamental physical dimensions comprising such 
Figure B.1 depicts hourly temperature  eries measured at 1.5 m 
variables. However, this has no adverse effect on the derivation of the 
aboveground by Viking Landers 1 and 2 (VL1 and VL2 respectively) in 
sought dimensionless π  products.
the late 1970s [60]. The VL1 record covers about half of a Martian year, 
i
Substituting the above roots in Eq. (A.2) reduces the original five 
while the VL2 series extends to nearly 1.6 years. The VL1 temperature 
unknowns to two: d and e, i.e.
series captures a summer-fall se son on a site located at about 1,500 m 
below Datum  levation in the subtropics of Mars’ Northern Hemisphere 
1
2e
−(3e+)
≈    P
  d
  e
ρ
(22.5o N). The arithmetic average of the series is 207.3 K (Fig. B.1a). 
s
r
x
r
x
          (A.5a)
Since the record lacks data from the cooler winter-spring season, this 
ACT 1982
These solution powers may now be assigned arbitrary values, although 
value is likely highe  than the actual mean annual temperature at that 
integers such as 0, 1 and -1 are preferable, for they offer the simplest 
ocation. Furtherm re, observations by the Hubble telescope from the 
solution leading to the construction of proper π  variables. Setting d = 0 
i
mid-1990s ind cated that the Red Planet may have cooled somewhat 
and e = -1 reduces Eq. (A.5a) to
since the time of the Viking mission [132,133]. Because of a thin 
1
2

3
1
atmosphere and the absence of significant cloud cover and perceptible 
T
T S
ρ −

(A.5b)
s
 
    x
water, temperature fluctuations near the surface of Mars are tightly 
providing the first pair of dimensionless products:
coupled to diurnal, seasonal and latitudinal variations in incident solar 
3
radiation. This causes sites located at the same latitude and equivalent 
π = T
P
;     π =
x
1
2
2
alti udes to have similar annual temperature means irrespective of 
T
ρ S
r
   
(A.6)
their longitudes [134]. Hence, one could reliably estimate a latitudinal 
The second pair of π  variables emerges upon setting d = -1 and e = 0 in 
temperature average on Mars using point observations from any 
i
Eq. (A.5a), i.e.
elevation by applying an appropriate lapse-rate correction for the 
average terrain elevation of said latitude. 
π = T
P
;     π = x
1
2
At 22.5o absolute latitude, the average elevation between Northern 
T
P
           (A.7)
r
r
and Southern Hemisphere on Mars is close to Datum level, i.e. about 
Thus, the original function (A.1) consisting of six dimensioned  1,500 m above the VL1 site. Adjusting the observed 207.3 K temperature 
variables has been reduced to a relationship between two dimensionless 
INFORMATION 
average at VL1 to Datum elevation using a typical near-surface Martian 
quantities, i.e. π  = f (π ). This relationship must further be investigated 
1
2
lapse rate of -4.3 K km-1 [78] produces ~201 K for the average summer-
through regression analysis  
fall temperature at that latitude. Since the mean surface temperature 
RELEASED UNDER THE 
 OFFICIAL 
Figure B.1: Near-surface hourly temperatures measured on Mars by (a) Viking Lander 1 at Chryse Planitia (22.48° N, 49.97° W, Elevation: -1,500 m); and (b) Viking 
Lander 2 at Utopia Planitia (47.97° N, 225.74° W, Elevation: -3,000 m) (Kemppinen et al. [60]; data downloaded from: http://www-k12.atmos.washington.edu/k12/
resources/mars_data-information/data.html). Black dashed lines mark the arithmetic average (T
) of each series. Grey dashed lines highlight the range of most 
mean
frequently reported GMAT values for Mars, i.e. 210–240 K. The average diurnal temperature can only exceed 210 K during the summer; hence, all Martian latitudes 
outside the Equator must have mean annual temperatures significantl  lower than 210 K.
Environ Pollut Climate Change, an open access journal 
Volume 1 • Issue 2 • 1000112


Citation: N kolov N, Zeller K (2017) New Insights on the Physical Nature of the Atmospheric Greenhouse Effect Deduced from an Empirical Planetary 
Temperature Model. Environ Pollut Climate Change 1: 112. 
Page 19 of 22
of a sphere is typical y lower than its subtropical temperature average, 
atmosphere. Figures 2 and 3 of Shirley et al. [136] depict nighttime 
we can safely conclude based on Figure B.1a that Mars’ GMAT is likely 
winter temperature profiles over the Mars’ northern and southern Polar 
below 201 K. The mean temperature at the VL2 site located at ~48o N 
Regions, respectively at about 75o absolute latitude. The average winter 
latitude and 3,000 m below Datum elevation is 191.1 K (Fig. B.1b). The 
surface temperature between the two Hemispheres for this latitude 
average terrain elevation between Northern and Southern Hemisphere 
is about 148.5 K. This compares favorably with 156.4 K produced by 
at 48o absolute latitude is about -1,500 m. Upon adjusting the VL2 
Eq. (B.1) for 75o (1.309 rad) latitude considering that MAT values are 
annual temperature mean to -1,500 m altitude using a lapse rate of 
expected to be higher than winter temperature averages. Figures 4 and 
-4.3 K km-1 we obtain 184.6 K. Since a planet’s GMAT numerical y fal s
5 of Shirley et al. [136] portray average temperature profiles retrieved 
between the mean temperature of the Equator and that of 42o absolute
by MGS-RST and MCS over lowlands (165o – 180o E) and highlands 
latitude, the above calculations suggest that Mars’ GMAT is likely
(240o - 270o E) of the Mars’ equatorial region (8o N - 8o S), respectively. 
between 184 K and 201 K.
For highlands (≈5 km above Datum), the near-surface temperature 
A close examination of the Viking record also reveals that average 
appears to be around 200 K, while for lowlands (≈2.5 km below Datum) 
diurnal temperatures above 210 K only occur on Mars during the 
it is ≈211 K. Since most of Mars’ equatorial region lies above Datum, it 
summer season and, therefore, cannot possibly represent an annual 
is likely that Mars’ equatorial MAT would be lower than 205.5 K and 
mean for any Martian latitude outside the Equator. On the other hand, 
close to our independent estimate of ≈203 K based on Curiosity Rover 
frequently reported values of Mars’ GMAT in excess of 210 K appear to 
measurements.
be based on the theoretical expectation that a planet’s average surface 
Mars’ GMAT () was calculated via integration of polynomial 
M
temperature should exceed the corresponding effective radiating  (B.1) using the formula:
temperature produced by Eq. (3) [6,78], which is  ≈ 212 K for Mars. 
π
e
2
This presumption is rooted in the a priori assumption that  represents 
 T L
L dL

          (B.2)
M
( )cos                                
e
a planet’s average surface temperature in the absence of atmospheric 
0
ACT 1982
greenhouse effect. However, Volokin and ReLlez [1] have shown  where 0 ≤ cosL ≤ 1 is a polar-coordinate area-weighting factor. 
that, due to Hölder’s inequality between integrals, the mean physical 
The result is  = 190.56 ± 0.7 K (Figure B.2). This estimate, while 
M
temperature of a spherical body with a tenuous atmosphere is always 
significantly lower than GMAT values quoted in recent publications, 
lower than its effective radiating temperature computed from the 
agrees quite well with spherical y integrated brightness temperatures 
global y integrated absorbed solar flux. In other words, Eq. (3) yield  
of Mars retrieved from remote microwave observations during the 
non-physical temperatures for spheres. Indeed, based on results from 
late 1960s and early 1970s [85-87]. Thus, according to Hobbs et al. 
a 3-D climate model Haberle  [130] concluded that Mars’ mean global 
[85] and Klein [86], the Martian mean global temperature (inferred
surface temperature is at least 8 K cooler than the planet’s effective 
from measurements at wavelengths between 1 and 21 cm) is 190 –
radiating temperature. Therefore, Mars’ GMAT must be inferred from 
193 K. Our  estimate is also consistent with the new mean surface
M
actual measurements rather than from theoretical calculations.
temperature of the Moon (197.35 K) derived by Volokin and ReLlez
In order to obtain a reliable estimate of Mars’ GMAT, we calculated 
[1] using output from a validated NASA thermo-physical model [29].
the mean annual temperatures at several Martian latitudes employing 
Since Mars receives 57% less solar ittadiance than the Moon and has
near-surface time series measured in-situ by Viking Landers and the 
a thin atmosphere that only delivers a weak greenhouse effect [9], it
Curiosity Rover, and remotely by the Mars Global Surveyor (MGS) 
makes a physical sense that the Red Planet would be on average cooler
spacecraft. The Radio Science Team (RST) at Stanford University 
than our Moon (i.e. < 197.3K). Moreover, if the average temperature 

utilized radio occultation of MGS refra tion data to retrieve seasonal 
time-series of near-surface atmosph ric temperature and pressure on 
INFORMATION 
Mars [61,62,135]. We utilized MGS-RST data obtained between 1999 
and 2005. Calculated mean temperatures from in-situ measurements 
RELEASED UNDER THE 
were adjusted to corresponding average  errain elevations of target 
latitudes using a lapse rate of -4.3 K km-1 [78]. Figure B.2 portrays 
the estimated Mean Annual n ar surface Temperatures (MAT) at five 
absolute Martian latitudes (gray dots) along with their standard errors 
(vertical bars). The equatorial MAT was calculated from Curiosity Rover 
observations; tempera ures at absolute latitudes 0.392 rad (22.48o) and 
0.837 rad (47.97o) were derived from VL measurements, while these 
at latitudes 1.117 rad (64o) and 1.396 rad (80o) were estimated from 
MGS-RST data. The black curve represents a third-order polynomial 
fitted through the latitudinal temperature averages and des
 OFFICIAL  cribed by the 
polynomial:
  (L)
2
3
= 202.888 − 0.781801 − 22.3673 − 3.16594             ( B. )
1
with  L being the absolute latitude (rad). MAT values predicted by 
Figure B.2:  Mean  annual  surface  air  temperatures  at  five Martian  absolute 
Eq. (B.1) for Mars’ Equatorial and Polar Regions agree well with 
latitudes (gray dots) es imated from data provided by Viking Landers, Curiosity 
Rover, and the Mars Global Surveyor Radio Science Team. Each dot represents 
independent near-surface temperatures remotely measured by the 
a  mean  annual  temperature  corresponding  to  the  average  terrain  elevation 
Mars Climate Sounder (MCS), a platform deployed after MGS in 
between  Northern  and  Southern  Hemisphere  for  particular  latitude. The  black 
curve  depicts  a  third-order  polynomial  (Eq.  B.1)  fitted through  the  latitudinal 
2006 [136]. Shirley et al. [136] showed that, although separated in 
temperature  means  using  a  non-linear  regression.  Mars’  GMAT,    =  190.56 
M
time by 2-5 years, MCS temperature profiles match quite well those 
K (marked by a horizontal gray dashed line) was calculated via integration of 
retrieved by MGS-RST especial y in the lower portion of the Martian 
polynomial (B.1) using formula (B.2). 
Environ Pollut Climate Change, an open access journal 
Volume 1 • Issue 2 • 1000112

Citation: N kolov N, Zeller K (2017) New Insights on the Physical Nature of the Atmospheric Greenhouse Effect Deduced from an Empirical Planetary 
Temperature Model. Environ Pollut Climate Change 1: 112. 
Page 20 of 22
of the lunar equator (Moon’s warmest latitude) is 213 K as revealed by 
21. Vignaux  GA,  Jain  S  (1988)  An  approximate  inventory  model  based  on 
NASA Diviner observations [1,29], it is unlikely that Mars’ mean global 
dimensional analysis. Asia-Pacifi  Journal of Operational Research 5: 117-123.
temperature would be equal to or higher than 213 K as assumed by 
22. Huntley HE (1967) Dimensional analysis. Dover Publications, New York.
many studies [6,76-78,80,131] 
23. Vignaux  GA  (1991)  Dimensional  analysis  in  data  modeling.  In:  Smith  CR, 
Published values of Mars’ average surface atmospheric pressures 
Erickson  G,  Neudorfer  PO  (eds)  Maximum  entropy  and  Bayesian  methods. 
Kluwer Academic Publishers, Seattle, pp 121-126 
range from 600 Pa to 700 Pa [6,32,78,80,83,84]. Since this interval was 
too broad for the target precision of our study, we employed MGS-RST 
24. Vignaux  GA,  Scott  JL  (1999)  Theory  and  methods:  Simplifying  regression
models using dimensional analysis. Aust N Z J Stat 41: 31–42.
data retrieved from multiple latitudes and seasons between 1999 and 
2005 to calculate a new mean surface air pressure for the Red Planet. 
25. Van Der Ha JC, Lappas VJ (2007) Long-term attitude drift of spinning spacecraft
Our analysis  produced P = 685.4 ± 14.2 Pa, an estimate within the 
under solar radiation torques. Journal of Guidance, Control and Dynamics 30: 
1470-1479.
range of previously reported values.
26. McMahon J, Scheeres D (2010) A new navigation force model for solar radiation
Funding Sources
pressure. Journal of Guidance, Control and Dynamics 33: 1418-1428.
This research did not receive any specific grant from funding agencies in the 
27. Buckingham E (1914) On physically similar systems: Illustrations of the use of
public, commercial, or not-for-profit sectors
dimensional equations. Phys Rev 4: 345-376.
References
28. Taylor FW (2010a) Planeta y a mospheres. Oxford University Press, USA.
1. Volokin D, ReLlez L (2014) On the average temperature of airless spherical 
29. Vasavada AR,  Bandfiel   JL,  Greenhagen  BT,  Hayne  PO,  Siegler  MA,  et  al. 
bodies and the magnitude of Earth’s atmospheric thermal effect. Springer Plus 
(2012) Lunar equa orial surface temperatures and regolith properties from the 
3: 723.
diviner lunar radiometer experiment. J Geophy Res 117: E00H18.
2. Hansen J, Johnson D, Lacis A, Lebedeff S, Lee P, et al. (1981) Climate impact
30. Blanco VM, McCuskey SW ( 961) Basic physics of the solar system. Addison-
of increasing atmospheric carbon dioxide. Science 213: 957-966.
Wesley, R ading MA.
ACT 1982
3. Peixoto JP, Oort AH (1992) Physics of climate. Springer-Verlag, New York.
31. Mölle  F (1964) Optics of  he lower atmosphere. Appl Optics 3: 157-166.
4. Wallace JM, Hobbs PV (2006) Atmospheric science: An Introductory survey, 
32. Will ams DR (2015) NASA Planetary Factsheet. NASA online.
Academic Press, California.
33. Beckenbach EF, Bellman R (1983) Inequalities. Springer Verlag, Berlin.
5. Lacis AA, Schmidt GA, Rind D, Ruedy RA (2010) Atmospheric CO : Principal
2
34. Rubincam DP (2004) Black body temperature, orbital elements, the Milankovitch 
control knob governing Ear h’s temperature. Science 330.
precession index and the Sever smith psychroterms. Theor Appl Climatol 79: 
6. Lacis AA, Hansen JE, Russell GL, Oinas V, Jonas J (2013) The role of long
111–131.
lived greenhouse gases as principal LW control knob that governs the global
3  Fix en DJ (2009) The temperature of the cosmic microwave background. Ap
surface temperature for past and future climate change. Tellus B 65: 19734.
J 707: 916.
7.  Schmidt GA, Ruedy R, Miller RL, Lacis AA et al. (2010) The attribution of  he
36. Paige DA, Foote MC, Greenhagen BT, Schofiel  JT, Calcutt S, et al. (2010) 
present-day total greenhouse effect. J Geophys Res 115: D20106
The lunar reconnaissance orbiter diviner lunar radiometer experiment. Space 
8. RamanathanV, Inamdar A (2006) The radiative forcing due to clouds and water 
Sci Rev 150: 125–160. 
vapor. Frontiers of Climate Modeling. Cambridge University Press, Cambridg , 
37. Leconte  J,  Forget  F,  Charnay  B,  Wordsworth  R,  Selsis  F,  et  al.  (2013)  3D 
pp. 119-151.
climate  modeling  of  close-in  land  planets:  circulation  patterns,  climate  moist 
9.  Pierrehumbert R (2010) Principles of planeta y climate. Cambridge University
bistability and habitability. Astron Astrophys 554, A69.
Press, New York.
38. Cengel  YA,  Turner  RH  (2004)  Fundamentals  of  thermal-flui   sciences. 
10. Pierrehumbert  R  (2011)  Infrared  radiat on  and  planetary  temperature.  Phys
McGraw-Hill, Boston.
Today 64: 33-38.
INFORMATION 
39. Steel RGD, Torrie J H (1960) Principles and procedures of statistics with special
11. Trenberth KE, Fasullo JT  Kiehl J (2009) Earth’s global energy budget. B Am 
reference to the biological sciences. McGraw Hill.
Meteorol Soc March: 311-323.
40. Zwillinger D (2003) Standard mathematical tables and formulae. Chapman &
RELEASED UNDER THE 
12. Stephens GL, Li1 J, Wild M, Clayson CA, Loeb N, et al. (2012) An update on 
Hall/CRC, p: 912.
Earth’s energy balance in light of the latest global observations. Nat Geosci 5: 
41. Atreya SK, Lorenz RD, Waite JH (2009) Volatile origin and cycles: Nitrogen 
691–696.
and methane. In: Brown RH, Lebreton JP, Waite JH (eds) Titan from Cassini-
13. Wild M, Folini D, Schär C, Loeb N, Dutton EG, et al. (2013) The global energy
Huygens. Springer, New York, pp: 177–200
balance from a surface perspective. Clim Dyn 40: 3107–3134.
42. Jennings DE, Flasar FM, Kunde VG, Samuelson RE, Pearl JC, et al. (2009) 
14. Wild M, Folini D, Hakuba MZ, Schar C, Seneviratne SI, et al. (2015) The energy
Titan’s surface brightness temperatures. Astrophys J 691: L103–L105.
balance over land and oceans: An assessment based on direct observations
43. Cottini V, Nixona CA, Jennings DE, et al. (2012) Spatial and temporal variations
and CMIP5 climate models. Clim Dyn 44: 3393.
in  Titan’s  surface  temperatures  from  Cassini  CIRS  observations.  Planetary
15. Bengtsson L, Bonnet R-M, Grinspoon D, Koumoutsaris D, Lebonnois S, et al. 
Space Sci 60: 62–71.
(2013) Towards understanding the climate of Venus: Applications of terrestrial 
 OFFICIAL 
44. Fegley  B,  Zolotov  MY,  Lodders  K  (1997)  The  oxidation  state  of  the  lower
models to our sister planet, ISSI scientifi  report series, Springer.
atmosphere and surface of Venus. Icarus 125: 416-439.
16. Rashevsky  N  (1960)  Mathematical  biophysics:  Physico-mathematical
45. Basilevsky AT,  Head  JW  (2003)  The  surface  of  Venus.  Rep  Prog  Phys  66: 
foundations of biology. Dover Publications, New York.
1699–1734.
17. Albertson ML, Barton JR, Simons DB (1961) Fluid mechanics for engineers. 
46. Mallama A,  Wang  D,  Howard  RA  (2006)  Venus  phase  function  and  forward
Prentice Hall, New Jersey.
scattering from H SO . Icarus 182: 10–22.
2
4
18. Yalin MS (1971) Theory of hydraulic models. MacMillan.
47. Basilevsky  AT,  McGill  GE  (2007)  Surface  evolution  of  Venus:  Esposito  LW 
19. Taylor ES (1974) Dimensional analysis for engineers. Clarendon Press, Oxford.
Exploring Venus as a Terrestrial Planet (Geophysical Monograph Series), Wiley 
176: 23-43.
20. Bender EA, (1978) An introduction to mathematical modeling. John Wiley and 
Sons, NY.
48. NASA Solar System Exploration (2014) Planets.
Environ Pollut Climate Change, an open access journal 
Volume 1 • Issue 2 • 1000112

Citation: N kolov N, Zeller K (2017) New Insights on the Physical Nature of the Atmospheric Greenhouse Effect Deduced from an Empirical Planetary 
Temperature Model. Environ Pollut Climate Change 1: 112. 
Page 21 of 22
49. Kopp G, Lean JL (2011) A new, lower value of total solar irradiance: Evidence
geometric albedo and bolometric Bond albedo of Neptune’s satellite Triton from 
and climate significanc . Geophys Res Lett 38: L01706.
Voyager observations. Geophys Res Lett 17: 1761–1764.
50. Jones PD, New M, Parker DE, Martin S, Rigor IG (1999) Surface air temperature
75. Elliot JL, Hammel HB, Wasserman LH, Franz OG, McDonald SW, et al. (1998) 
and its changes over the past 150 years, Rev Geophys 37: 173–199.
Global warming on Triton. Nature 393: 765-767.
51. NOAA  National  Climatic  Data  Center  (2014)  Global  surface  temperature
76. Smil V (2003) The Earth’s biosphere: Evolution, dynamics and change. MIT 
anomalies. Online publication by NOAA.
Press.
52. Smith TM, Reynolds RW, Peterson TC, Lawrimore J, et al. (2008) Improvements
77. Vázquez  M.  Hanslmeier  A  (2006)  Ultraviolet  radiation  in  the  solar  system. 
to  NOAA’s  Historical  Merged  Land-Ocean  Surface  Temperature  Analysis
Springer, The Netherlands.
(1880-2006). J Climate 21: 2283-2296.
78. Barlow N (2008) Mars: An introduction to its interior, surface and atmosphere. 
53. Trenberth KE, Smith L (2005) The mass of the atmosphere: A constraint on
Cambridge University Press, Cambridge.
global analyses. J Climate 18: 864-875.
79. Rapp D (2008) Human missions to Mars: Enabling technologie  for exploring
54. Stephens GL, O’Brien  D, Webster PJ, Pilewski P, Kato S, et al. (2015) The 
the red planet. Springer, Germany.
albedo of Earth. Rev Geophys 53.
80. Taylor  FW  (2010b)  The  scientifi   exploration  of  Mars   Cambridge  University
55. Loeb NG, Wielicki BA, Doelling DR, Smith GL, Keyes DF, et al. (2009) Toward 
Press, Cambridge NY
optimal closure of the Earth’s top-of-atmosphere radiation budget. J Climate
22:748–766.
81. Lissauer JJ, Pater I (2013) Fundamental planetary science: Physics, chemistry
and habitability. Cambridge University Press, New York NY.
56. Lucey,  P  Korotev  RL,  Gillis  JJ,  Taylor  LA,  Lawrence  D,  et  al.  (2006) 
82. Fenton  LK,  Geis ler  PE,  Haberle  RM  (2007)  Global  warming  and  climate
Understanding  the  lunar  surface  and  space-Moon  interactions.  Reviews  in 
forcing by recent a bedo changes on Mars. Nature 446: 647-649.
Mineralogy and Geochemistry 60: 83–219.
83. Jakosky BM  Phillips RJ (2001) Mars’ vo atile and climate history. Nature 412: 
57. Keihm SJ (1984) Interpretation of the lunar microwave brightness temperature
237-244
spectrum: feasibility of orbital heat flo  mapping. Icarus 60: 568–589.
84. Catling DC  Leovy C (2007) Ma s atmosphere: History and surface interactions. 
ACT 1982
58. Vasavada  AR,  Paige  DA,  Wood  SE  (1999)  Near-surface  temperatures  on
In: Encyclopedia of the Solar System. Academic Press, pp: 301-314.
Mercury and the Moon and the stability of polar ice deposits. Icarus 141: 179–
193.
85. Hobbs RW, McCullough TP, Waak JA (1968) Measurements of Mars at 1.55 cm
and 0.95 cm wavelengths. Icarus 9: 360–363.
59. Mat hews G (2008) Celestial body irradiance determination from an underfille
satellite radiometer: Application to albedo and thermal emission measurements
86  Klein MJ (1971) Mars: measurements of its brightness temperature at 1.85 and
of the Moon using CERES. Applied Optics 47: 4981-4993.
3.75 cm wave ength. Icarus 14: 210-213.
60. Kemppinen O, Tillman JE, Schmidt W, Harri AM (2013) New analysis software 
87. Briggs  FH,  Drake  FD  (1972)  Interferometric  observations  of  Mars  at  21  cm
for  Viking  Lander  meteorological  data.  Geosci  Instrum  Me hod  Data  Syst  2: 
wavel ngth. Icarus 17: 543-547.
61–69.
88. Van man D, Reedy R, Heiken G, Haskin L, Greive R, et al. (1991) The lunar
61. Hinson  DP,  Smith  MD,  Conrath  BJ  (2004)  Comparison  of  atmospheric 
environment. In: Lunar Sourcebook: A User’s Guide to  he Moon. Cambridge
temperatures obtained through infrared sounding and radio occultation by Ma s
University Press, Cambridge, pp 27-60
Global Surveyor. J Geophys Res 109: E12002.
89. Morice CP, Kennedy JJ, Rayner NA, Jones PD (2012) Quantifying uncertainties 
62. Hinson DP (2006) Radio occultation measurements of tran ient eddies in  he
in global and regional temperature change using an ensemble of observational
northern hemisphere of Mars. J Geophys Res 11 : E05002.
estimates: The HadCRUT4 data set. J Geophys Res 117: D08101.
63. Bandfiel  JL, Wolff MJ, Smith MD, Daniel JM (2013) Radiome ric comparison 
90. Peterson TC, Vose RS (1997) An overview of the global historical climatology
of Mars Climate Sounder and Thermal Emis ion spectrometer measurements. 
network temperature database. Bull Am Meteorol Soc 78: 2837-2849
Icarus 225: 28-39.
91. Smith  TM,  Reynolds  RW  (2005) A  global  merged  land  air  and  sea  surface
64. Younkin RL (1974) The albedo of Titan. Ic rus 21: 219–229.
temperature  reconstruction  based  on  historical  observations  (1880-1997).  J 
Climate 18: 2021-2036.
INFORMATION 
65. Hanel  RA,  Pearl  JC,  Samuelson  RE  (1985)  The  bolometric  bond  albedo  of
Titan. Bulletin of the Astronomical Society 17: 739.
92. Hansen J, Sato M, Russell G, Kharecha P (2013) Climate sensitivity, sea level
and atmospheric carbon dioxide. Phil Trans R Soc A 371: 20120294.
66. Neff  JS,  Ellis  TA,  Apt  J,  Bergstralh  JT  (1985)  Bolometric  albedos  of  Titan, 
RELEASED UNDER THE 
Uranus, and Neptune. Icarus 62: 425–432.
93. Gladstone GR, Stern SA, Ennico K, Olkin CB, Weaver HA, et al. (2016) The 
atmosphere of Pluto as observed by New Horizons. Science 351.
67. Fulchignoni  M,  Ferri  F,  Angrilli  F,  Ball  A ,  Nub-Bar  A,  et  al.  (2005)  In  situ 
measurements of the physical characte istics of Titan’s environment. Nature
94. NASA JPL Voyager Mission (2013) Titan. Online publica ion by the California
438: 785-791.
Institute of Technology.
68. Niemann HB, Atreya SK, Bauer SJ, Carignan GR, Demick JE, et al. (2005) The 
95. NASA  JPL  Ephemeris  (2014)  Horizons  web  interface:  California  Institute  of 
abundances of constituents of Titan’s atmosphere from the GCMS instrument
Technology.
on the Huygens probe  Nature 438: 779–784.
96. Stolk  H,  Gates  K,  Hanan  J  (2003)  Discovery  of  emergent  natural  laws
69. Griffit  CA (2007) Titan’s lower atmosphere. AIP Conf Proc 930: 3-36.
by  hierarchical  multi-agent  systems.  In:  Proceedings  of  the  International 
Conference  on  Intelligent  Agent  Technology.  IEEE/WIC  International 
70. Mitri  G,  Showmana AP,  Lunine  JI,  Lorenz  RD  (2007)  Hydrocarbon  lakes  on 
Conference on Intelligent Agent Technology, Halifax, Canada, pp: 75-82.
 OFFICIAL 
Titan. Icarus 186: 385–394.
97. Wild M (2009) Global dimming and brightening: A review. J Geophys Res 114: 
71. Li L, Nixon CA, Achterberg RK, Smith MA, Gorius NJP, et al. (2011) The global
D00D16.
energy balance of Titan. Geophys Res Lett 38: L23201.
98. Herman  J,  DeLand  MT,  Huang  L-K,  Labow  G,  Larko  D,  et  al (2013)  A  net 
72. Schinder  PJ,  Flasar  FM,  Marouf  EA,  French  RG,  Mcghee-French  C,  et  al. 
decrease in the Earth’s cloud, aerosol, and surface 340 nm reflectivit  during
(2012)  The  structure  of  Titan’s  atmosphere  from  Cassini  radio  occultation: 
the past 33 years (1979–2011). Atmos Chem Phys 13: 8505–8524.
Occultation from the Prime and Equinox missions. Icarus 221: 1020–1031.
99. Stanhill GO, Rosa AR, Cohen S, Achiman O (2014) The cause of solar dimming
73. Lellouch E, de Bergh C, Sicardy B, Ferron S, Kaufl HU (2010) Detection of 
and brightening at the Earth’s surface during  he last half century: Evidence from
CO in Triton’s atmosphere and the nature of surface-atmosphere interac ions. 
measurements of sunshine dura ion J Geophys Res Atmos 119: 10902–10911.
Astron Astrophys 512: L8.
100. Snyder CW (2016) Evolution of global temperature over the past two million
74. Nelson RM, Burattl BJ, Wallis BD, Smythe WD, Horn LJ, et al. (1990) Spectral
years. Nature 538: 226–228.
Environ Pollut Climate Change, an open access journal 
Volume 1 • Issue 2 • 1000112

Citation: N kolov N, Zeller K (2017) New Insights on the Physical Nature of the Atmospheric Greenhouse Effect Deduced from an Empirical Planetary 
Temperature Model. Environ Pollut Climate Change 1: 112. 
Page 22 of 22
101. Heinemann  M,  Timmermann  A,  Timm  OE,  Saito  F,  Abe-Ouchi  A  (2014) 
120. Stern SA (2008) The New Horizons Pluto Kuiper belt mission: An overview
Deglacial ice sheet meltdown: orbital pace making and CO  effects. Clim Past
with historical context. Sp Sci Rev 140: 3-21.
2
10: 1567–1579.
121. Pappalardo RT, Vance S, Bagenal F, Bills BG, Blaney DL, et al. (2013) Science
102. Stap  LB,  van  de  Wal  RSW,  de  Boer  B,  Bintanja  R,  Lourens  LJ  (2014) 
potential from a Europa lander. Astrobiology 13: 740-773.
Interaction of ice sheets and climate during the past 800 000 years. Clim Past 
10: 2135–2152.
122. Olkin CB, Young LA, Borncamp D, Pickles A, Sicardy B, et al. (2013) Pluto’s 
atmosphere does not collapse. Cornell University Library, USA.
103. McGehee R, Lehman C (2012) A paleoclimate model of ice-albedo feedback
forced by variations in Earth’s orbit. Siam J App Dyn Sys 11: 684–707.
123. Mallama A, Wang D, Howard RA (2002) Photometry of Mercury from SOHO/
LASCO and Earth. Icarus 155: 253–264.
104. Doughty AM, Schaefer JM, Putnam AE, Denton GH, Kaplan MR, et al. (2015) 
Mismatch of glacier extent and summer insolation in Southern Hemisphere
124. Spencer  JR,  Tamppari  LK,  Martin  TZ,  Travis  LD  (1999)  Temperatures
mid-latitudes. Geology G36477.1.
on  Europa  from  Galileo  photopolarimeter-radiometer:  Nighttime  thermal
anomalies. Science 284: 1514-1516.
105. Maslin MA, Brierley CM (2015) The role of orbital forcing in the early middle
125. Moore JM, Chapman CR, Bierhaus EB, Greeley R, Chuang FC  et al. (2004) 
Pleistocene transition. Quat Int 389: 47–55
Callisto.  In:  Jupiter:  The  planet,  satellites  and  magnetosphere.  Cambridge
106. Berner  RA  (2006)  Geological  nitrogen  cycle  and  atmospheric  N   over
University Press, pp 397-426
2
Phanerozoic  ime. Geology 34: 413-415.
126. Veeder GJ, Matson DL, J hnson TV, Blaney DL  Goguen JD (1994) Io’s heat
107. Vladilo  G,  Murante  G,  Silva  L,  Provenzale  A,  Ferri  G,  et  al.  (2013)  The
flo  from infrared radiometry: 1983–1993. J Geophys Res 99: 17095–17162.
habitable  zone  of  Earth-like  planets  with  different  levels  of  atmospheric
127. Sotin  C,  Head  II   JW   Tobie  G  (2002)  Europa:  Tidal  heating  of  upwelling
pressure. Astrophys J 767: 65.
thermal plumes and the origin of lenticulae and chaos melting. Geophys Res
108. Svensmark  H,  Bondo  T,  Svensmark  J  (2009)  Cosmic  ray  decreases  affect
Lett 29: 74-1–74-4. 
atmospheric aerosols and clouds. Geophys Res Lett 36: L15101.
128. Stern SA, Bagenal F, Ennico K  Gladstone GR, Grundy WM, et al. (2015) The 
109. Svensmark  J,  Enghoff  MB,  Svensmark  H  (2012)  Effects  of  cosmic  ray
Pluto system: initial results from its exploration by New Horizons. Science 350.
ACT 1982
decreases on cloud microphysics. Atmos Chem Phys Discuss 12: 3595–3617.
129. Asadi A, Hassan MM (2014) E aluation of the thermal performance of a roof-
110. Svensmark J, Enghoff MB, Shaviv N, Svensmark H (2016) The response of 
mo nted radiant b rrier in residential buildings: Experimental study. J Building
Phys 38: 66–80
clouds and aerosols to cosmic ray decreases. J Geophys Res Space Physics 
121: 8152-8181.
130  Haberle RM (2013) Estimating the power of Mars’ greenhouse effect. Icarus
223: 619-620.
111. Voiculescu M, Usoskin I, Condurache-Bota S (2013) Clouds blown by the solar
wind. Environ Res Lett 8: 045032.
131. Schulze Makuch D, Méndez A, Fairén AG, von Paris P, Turse C, et al. (2011) 
A two-tiered approach to assessing the habitability of exoplanets. Astrobiology 
112. Scafetta N (2016) High resolu ion coherence analysis between planetary and 
11: 1041–1052.
climate oscillations. Adv Space Res 57: 2121–2135.
132. Savage D, Jones T, Villard R (1995) Hubble monitors weather on neighboring
113. Kirkby J, Duplissy J, Sengupta K, Frege C, Gordon H, et al. (2016) Ion induced
planets. Hubble News Release Archive.
nucleation of pure biogenic particles. Nature 533: 521–526.
133. Clancy RT, Grossman AW, Wolff MJ, James PB, Rudy DJ, et al. (1996) Water
114. Palle E, Goode PR, Montañés-Rodríguez P, Shumko A  Gonzalez-Merino B, 
vapor saturation at low altitudes around Mars aphelion: A key to Mars climate? 
et al. (2016) Earth’s albedo variations 1998–2014 as measured from ground-
Icarus 122: 36–62.
based earthshine observations. Geophys Res Lett 43: 4531–4538.
134. Wilson RJ, Richardson MI (2000) The Martian atmosphere during the Viking 
115. Raulin F (2008) Planetary science: Organic lakes on Titan. Nature 454: 587-
mission:  I  Infrared  measurements  of  atmospheric  temperatures  revisited. 
589.
Icarus 145: 555–579.
116. Sharma  P,  Byrne  S  (2011)  Comparison  of  Titan’s  north  polar  lakes  with
135. Mars Global Surveyor Radio Science Team (2007) The Daily Martian Weather
terrestrial analogs. Geophys Res Lett 38  L24203.
Report. Online publication by Stanford University, USA.
INFORMATION 
117. Goldblatt  C,  Watson  AJ  (20 2)  The  runaway  greenhouse:  Implications  for 
136. Shirley  JH,  Schofiel   J,  Kleinböhl  A,  Abbatt  JPD,  Lollar  BS,  et  al.  (2011) 
future climate change, geoengineering and planetary atmospheres. Phil Trans
Comparison  of  MGS  Radio  Science  Mean  temperature  profile   with  Mars 
R Soc A 370: 4197–4216.
Climate  Sounder  (MCS)  results.  In:  Forget  CF  et  al.  (eds),  ‘The  Fourth
RELEASED UNDER THE 
International Workshop on the Mars Atmosphere: Modelling and observation’, 
118. Kasting  JF  (1988)  Runaway  and  moist  greenhouse  atmospheres  and  the
Paris, France.
evolution of Earth and Venus. Icarus 74:  72-494.
119. McClintock WE, Lankton MR (2007) The Mercury atmospheric and surface
composition spectrometer for the MESSENGER mission. Space Sci Rev 131: 
481-521.
 OFFICIAL 
Citation:  Nikolov  N,  Zeller  K  (2017)  New  Insights  on  the  Physical  Nature  of 
the  Atmospheric  Greenhouse  Effect  Deduced  from  an  Empirical  Planetary 
Temperature Model. Environ Pollut Climate Change 1: 112. 
Environ Pollut Climate Change, an open access journal 
Volume 1 • Issue 2 • 1000112


Local Government New Zealand leads on global warming
nzcpr.com/local-government-new-zealand-leads-on-global-warming/
Bryan
Leyland
Posted on July 1, 2018 By Bryan Leyland
Local Government New Zealand have embarked on a “Climate Change Project” focused on
adapting and mitigating “climate change” – properly described as man-made global
warming.
When faced with a potential risk, the rational approach is to make sure that the risk is real,
assess its magnitude, decide if anything needs to be done, and if so, what is the cheapest
and most effective solution. 
In spite of the fact that no one has any convincing evidence based on observations that
man-made global warming real and dangerous LGNZ have jumped to the conclusion that
ACT 1982
the risk is real, urgent action is needed and lots of our money and resources must be spent
on “fighting climate change”. Taking an objective look at al  the evidence never even crossed
their minds.
If they had looked at the evidence, they would have got a big surprise. 
They would have discovered that world temperatures have increased by about half the
predicted amount over the last 20 years and New Zealand has hardly warmed it al . This
would – or should – tel  them that the computer models which the climate scientists rely
upon for predicting future climate are worthless. There is nothing abnormal about the
modest amount of warming that has occurred as we recover from the Little Ice Age. 
INFORMATION 
They would also discover that sea level rise in New Zealand – and the rest of the world – has
RELEASED UNDER THE 
been steady at between 1 5 and 2 mm per year for the last hundred years and shows no
sign of the claimed recent rapid increase. They would also discover that there is no reason –
other than the failed climate models – to assume that it wil  rise more rapidly in the future.
If they studied storms, floods and droughts in New Zealand and the rest of the world they
would find that recent weather is rather better than it was in the past. The IPCC agrees.
 OFFICIAL 
If they looked at the history of atol  formation they would realise that coral atol s were able
to keep up with a sea level rise of 3000 mm per century at the end of the ice age. It fol ows
that they cannot be in danger from the current tiny rate of sea level rise. Pacific islands do
have real problems, but they are not caused by sea level rise.
1/2

If they looked further they would discover that there are many very credible papers based
on observations and experiments that indicate a very high probability that the world wil
soon enter a cooling cycle. Right now sunspot levels are lower than they have been since the
Little Ice Age and the correlation between sunspot levels and temperatures is very strong. 
A Danish professor has established a cause and effect relationship between sunspot cycles,
cosmic rays, low clouds and global temperatures. When sunspot levels are low, the
magnetic shield emitted by the sun is low and this al ows more high energy cosmic rays to
reach lower levels in the atmosphere. When they do, they cause condensation and this
triggers cloud formation. Other scientists have analysed past climate cycles and concluded
that there is a high risk of global cooling.
While they regard carbon dioxide as a dangerous pol utant, without it, life on earth could not
exist. The reality is that it is essential to life and plant growth and the recent rise in
concentration has increased agricultural productivity by about 15%. A big win for New
Zealand’s economy.  
ACT 1982
They might also be interested to discover that neither the United Nations Intergovernmental
Panel on Climate Change, the Royal Society of New Zealand nor Prof Jim Renwick can
provide convincing evidence based on observations of the real world that man-made
greenhouse gases cause dangerous global warming. The evidence simply does not exist.
Until this evidence is discovered – if it ever is – the only rational conclusion is that man-
made global warming is, in al  probability  the biggest hoax in the history of the world. 
It is tragic that Local Government New Zealand have bought into the global warming hoax. 
We should not be squandering our money and damaging our economy in a futile attempt to
solve a problem that  according to the evidence, does not exist.
INFORMATION 
RELEASED UNDER THE 
 OFFICIAL 
2/2

Document Outline