ArticlePDF Available

Abstract and Figures

Using data series on atmospheric carbon dioxide and global temperatures we investigate the phase relation (leads/lags) between these for the period January 1980 to December 2011. Ice cores show atmospheric CO2 variations to lag behind atmospheric temperature changes on a century to millennium scale, but modern temperature is expected to lag changes in atmospheric CO2, as the atmospheric temperature increase since about 1975 generally is assumed to be caused by the modern increase in CO2. In our analysis we use eight well-known datasets: 1) globally averaged well-mixed marine boundary layer CO2 data, 2) HadCRUT3 surface air temperature data, 3) GISS surface air temperature data, 4) NCDC surface air temperature data, 5) HadSST2 sea surface data, 6) UAH lower troposphere temperature data series, 7) CDIAC data on release of anthropogene CO2, and 8) GWP data on volcanic eruptions. Annual cycles are present in all datasets except 7) and 8), and to remove the influence of these we analyze 12-month averaged data. We find a high degree of co-variation between all data series except 7) and 8), but with changes in CO2 always lagging changes in temperature. The maximum positive correlation between CO2 and temperature is found for CO2 lagging 11–12 months in relation to global sea surface temperature, 9.5–10 months to global surface air temperature, and about 9 months to global lower troposphere temperature. The correlation between changes in ocean temperatures and atmospheric CO2 is high, but do not explain all observed changes.
Content may be subject to copyright.
(This is a sample cover image for this issue. The actual cover is not yet available at this time.)
This article appeared in a journal published by Elsevier. The attached
copy is furnished to the author for internal non-commercial research
and education use, including for instruction at the authors institution
and sharing with colleagues.
Other uses, including reproduction and distribution, or selling or
licensing copies, or posting to personal, institutional or third party
websites are prohibited.
In most cases authors are permitted to post their version of the
article (e.g. in Word or Tex form) to their personal website or
institutional repository. Authors requiring further information
regarding Elsevier’s archiving and manuscript policies are
encouraged to visit:
http://www.elsevier.com/copyright
Author's personal copy
The phase relation between atmospheric carbon dioxide and global temperature
Ole Humlum
a,b,
, Kjell Stordahl
c
, Jan-Erik Solheim
d
a
Department of Geosciences, University of Oslo, P.O. Box 1047 Blindern, N-0316 Oslo, Norway
b
Department of Geology, University Centre in Svalbard (UNIS), P.O. Box 156, N-9171 Longyearbyen, Svalbard, Norway
c
Telenor Norway, Finance, N-1331 Fornebu, Norway
d
Department of Physics and Technology, University of Tromsø, N-9037 Tromsø, Norway
abstractarticle info
Article history:
Received 25 April 2012
Accepted 25 August 2012
Available online 30 August 2012
Keywords:
Carbon dioxide
Global temperature
Sea surface temperature
Volcanic eruptions
Using data series on atmospheric carbon dioxide and global temperatures we investigate the phase relation
(leads/lags) between these for the period January 1980 to December 2011. Ice cores show atmospheric CO
2
var-
iations to lag behind atmospheric temperature changes on a century to millennium scale, but modern tempera-
ture is expected to lag changes in atmospheric CO
2
, as the atmospheric temperature increase since about 1975
generally is assumed to be caused by the modern increase in CO
2
. In our analysis we use eight well-known
datasets: 1) globally averaged well-mixed marine boundary layer CO
2
data, 2) HadCRUT3 surface air tempera-
ture data, 3) GISS surface air temperature data, 4) NCDC surface air temperature data, 5) HadSST2 sea surface
data, 6) UAH lower troposphere temperature data series, 7) CDIAC data on release of anthropogene CO
2
,and
8) GWP data on volcanic eruptions. Annual cycles are present in all datasets except 7) and 8), and to remove
the inuence of these we analyze 12-month averaged data. We nd a high degree of co-variation between all
data series except 7) and 8), but with changesin CO
2
always lagging changes in temperature.The maximum pos-
itive correlation between CO
2
and temperature is found for CO
2
lagging 1112 months in relation to global sea
surface temperature, 9.510 months to global surface air temperature, and about 9 months to global lower tro-
pospheretemperature. The correlation betweenchanges in ocean temperatures and atmospheric CO
2
is high, but
do not explain all observed changes.
© 2012 Elsevier B.V. All rights reserved.
1. Introduction
IcecorerecordsindicatethatthegreenhousegasCO
2
co-varied
with the global temperature over several glacialinterglacial cycles,
suggesting a link between natural atmospheric greenhouse gas var-
iations and temperature on long time scales (IPCC AR4, 2007; Lüthi
et al., 2008).
Over the last 420 kyr variations in atmospheric CO
2
broadly
followed temperature according to ice cores, with a typical delay of
several centuries to more than a millennium (Lorius et al., 1990;
Mudelsee, 2001; Caillon et al., 2003). Atmospheric CO
2
is therefore
not initiating the large glacialinterglacial climate changes, and pre-
sumably these are controlled by orbital Milankovitch cycles. It has
however been suggested that the subsequent CO
2
-rise may amplify
or even in certain periods precede the global temperature increase
initiated by Milankovitch cycles, but the interpretation of the proxy
data is ambiguous with regard to this (Alley and Clarck, 1999;
Shackleton, 2000; Toggweiler and Lea, 2010; Shakun et al., 2012).
The observed time lag between atmospheric temperature and CO
2
from ice cores is thought to be caused by the slow vertical mixing that
occurs in the oceans, in association with the decrease in the solubility
of CO
2
in ocean water, as its temperature slowly increase at the end of
glacial periods (Martin et al., 2005), leading to subsequent net
out-gassing of CO
2
from the oceans (Toggweiler, 1999).
Direct measurements of temperatures and atmospheric CO
2
with
good time resolution are essential to understand empirically the effects
of CO
2
on modern global temperature changes. The rst in situ continuous
measurements of atmospheric CO
2
made by a high-precision non-
dispersive infrared gas analyzer were implemented by C.D. Keeling from
the Scripps Institution of Oceanography (SIO). These measurements
were initiated in 1958 at Mauna Loa, Hawaii, located at 19°N in the Pacic
Ocean (Keeling et al., 1995). These data documented that not only was the
amount of CO
2
increasing in the atmosphere since 1958, but also that the
rise was modulated by annual cycles caused by seasonal changes in ocean
surface temperature and photosynthesis in the terrestrial biosphere.
The Mauna Loa measurements were followed by other continuous in
situ measurements at a limited number of other observation sites in both
hemispheres (Conway et al., 1994; Nakazawa et al., 1997; Langenfelds
et al., 2002). In the 1980s and 1990s, however, it was recognized that
a greater coverage of CO
2
measurements over continental areas was
required to provide the basis for estimating sources and sinks of
Global and Planetary Change 100 (2013) 5169
Corresponding author at: Department of Geosciences, University of Oslo, P.O. Box
1047 Blindern, N-0316 Oslo, Norway. Tel.: + 47 41403157.
E-mail address: Ole.Humlum@geo.uio.no (O. Humlum).
0921-8181/$ see front matter © 2012 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.gloplacha.2012.08.008
Contents lists available at SciVerse ScienceDirect
Global and Planetary Change
journal homepage: www.elsevier.com/locate/gloplacha
Author's personal copy
atmospheric CO
2
over land as well as oceanregions, and to provide basis
for calculating a good estimate of theglobal amount of atmospheric CO
2
.
2. Modern global CO
2
and temperature
Today, an extensive network of international air sampling sites is
operated by the National Oceanic and Atmospheric Administration's
Global Monitoring Division (NOAA/GMD) in the USA. This organiza-
tion has measured carbon dioxide and other greenhouse gases for
several decades at a globally distributed network of air sampling
sites (Conway et al., 1994; IPCC AR4, 2007). A global average is
constructed by tting a smoothed curve as a function of time to
each site, and then the smoothed value for each site is plotted as a
function of latitude for 48 equal time steps per year (IPCC AR4,
2007). A global average is calculated from the latitude plot at each
time step (Masarie and Tans, 1995), based on measurements from a
subset of network sites. Only sites where samples are predominantly
of well-mixed marine boundary layer (MBL) air representative of a
large volume of the atmosphere are considered for the global CO
2
data series (IPCC AR4, 2007). These key sites are typically at remote
marine sea level locations with prevailing onshore winds, to mini-
mize the effects of inland vegetation and industries. Measurements
from sites at higher altitude and from sites close to anthropogenic
and natural sources and sinks are excluded from the global CO
2
esti-
mate. The MBL data provide a low-noise representation of the global
trend and allows making the estimate directly from the data without
the need for applying an atmospheric transport model (IPCC AR4,
2007).
Global monthly CO
2
data (NOAA) are available from January 1980,
and is shown graphically in Fig. 1, along with the monthly global sea
surface temperature (HadSST2) and the monthly global surface air
temperature (HadCRUT3), using data published by the University of
East Anglia and the Hadley Centre, UK. In addition, in the present
study we also analyze global air temperature data from the Goddard
Institute for Space Studies (GISS) in USA, the National Climatic Data
Center (NCDC) in USA, and lower troposphere temperature data pub-
lished by the University of Alabama (UAH), Huntsville, USA. At the
end of the paper a list of URL's used to obtain the data used can be
found. All these monthly data series are now sufciently long to
have collected a population of climate perturbations, and they are
therefore likely to reveal essentials of the coupling between atmo-
spheric CO
2
and temperature in modern time.
Global atmospheric CO
2
(Fig. 1) has increased steadily during the
entire observation period since 1980. There is, however, a pronounced
annual cyclic variation superimposed on this overall development,
caused by seasonal changes in the magnitude of sources and sinks for
CO
2
, controlled by dynamic exchanges with oceans and vegetation
(IPCC AR4, 2007). The two temperature series HadSST2 and HadCRUT3
also show an overall increase over the period, but their detailed devel-
opment is more complex than for CO
2
. In addition, they only show
small net changes since early 2002 (Scafetta, 2011).
In general, the two temperature records are seen to vary in close
concert with each other. This also applies for the three other temper-
ature records considered in this study (GISS, NCDC and UAH), but
these are not displayed in Fig. 1 to avoid visual congestion. All four
temperature records display rhythmic annual variations because of
the uneven hemispherical distribution of land and ocean, although
this is not readily apparent from the diagram, as other short-term
variations tend to dominate.
Resolving the degree of coupling between CO
2
and temperature is
not visually straightforward as illustrated by Fig. 1, but obviously re-
quires a more elaborate approach to the data series.
3. DIFF12 values
Before analyzing the monthly data, being interested in longer than
annual variations, we rst removed the annual cycle from the global
atmospheric CO
2
data series by calculating a 12-month running aver-
age. This implies that we here consider the annual variation as noise
only, and instead are looking for the underlying longer signal, the
overall CO
2
increase. As the signal tends to be almost the same from
one monthly observation to nearby observations and the noise does
not, an average of several adjacent monthly observations will tend
to converge on the value of the signal alone. The most serious
Fig. 1. Monthly global atmospheric CO
2
(NOOA; green), monthly globalsea surface temperature (HadSST2; blue stippled) and monthly global surface airtemperature (HadCRUT3; red),
since January 1980. Last month shown is December 2011. (For interpretation of the references to color in this gure legend, the reader is referred to the web version of the article.)
52 O. Humlum et al. / Global and Planetary Change 100 (2013) 5169
Author's personal copy
consequence of smoothing or ltering is the shift of peaks and
troughs in the smoothed curve, relative to the original data. If several
data series are to be compared, identical ltering must therefore be
applied at all series, as spurious effects else may arise, perhaps even
inviting a false interpretation (see, e.g., discussion in Stauning, 2011).
We next calculated the difference between the average CO
2
con-
centration for the last 12 months and the average for the preceding
12 month period, referred to as DIFF12 in the following (see Fig. 1
for graphical explanation). In other words, DIFF 12 represents the
yearly net change with monthly time resolution. Technically DIFF12
values were plotted for the last month considered in the calculation.
By this we can efciently visualize and analyze changes in the overall
atmospheric CO
2
increase since January 1980, without confounding
inuence by the annual variation.
It is important to stress that the presence of a DIFF12 peak do not
indicate the presence of a concentration peak in the original CO
2
data, but instead a period with more rapid increase than else. It is
also worth emphasizing that by studying DIFF12 values we are inves-
tigating change rates in the amount of atmospheric CO
2
, and not the
total amount of CO
2
itself. However, by integrating DIFF12 values
over the observation period, changes in the total amount of CO
2
are
of cause addressed. Thus, to the degree that DIFF12 values can be
explained, changes in the total amount of atmospheric CO
2
are there-
fore also explained.
We next compared the result of the DIFF12 CO
2
calculation with
similar DIFF12 values for the global sea surface (HadSST2) and the
global surface air temperature (HadCRUT3). By this all monthly data
series were exposed to identical operations and the results are there-
fore comparable. To ensure that no spurious effects were introduced
by the data ltering we carried out a similar analysis on the unltered
monthly data, by calculating DIFF1
annual
,dened as the difference be-
tween 1 month and the identical month 1 yr before, e.g. January 2000
minus January 1999, etc. Finally, all DIFF1
annual
and DIFF12 series
were calculated and plotted in a time-change diagram (Fig. 2), to ob-
tain graphical overview.
From Fig. 2 it is seen that the 12-month ltering process results in
a displacement of DIFF12 peak and minimum values 12 months
ahead in time in relation to the unltered DIFF1
annual
values, exactly
as expected from denition of DIFF12. More important, however, it
is also seen that the relation between peaks and lows in the different
data series as to their relative timing remains unchanged. If no time
displacement of peaks and lows by ltering was important for the fol-
lowing analysis, this might have been obtained simply by changing
the plotting convention, and plotting DIFF12 at the last month in
the rst 12-month interval, instead of at the last month in the second
interval. Thus, by using a 12-month lter as outlined above we ac-
quire a considerably clearer picture of the underlying signal in the in-
dividual data series (Fig. 2), compared to the unltered data (Fig. 1).
Fig. 2. 12-month change of global atmospheric CO
2
concentration (NOAA; green), global sea surface temperature (HadSST2; blue) and global surface air temperature (HadCRUT3;
red dotted). The upper panel shows unltered monthly values (e.g. January 2000 minus January 1999), while the lower panel shows ltered values (DIFF12, the difference between
the average of the last 12 month and the average for the previous 12 months for each data series). The numbers (19) on DIFF12 CO
2
peaks and the thin white lines refer to Table 1.
(For interpretation of the references to color in this gure legend, the reader is referred to the web version of the article.)
53O. Humlum et al. / Global and Planetary Change 100 (2013) 5169
Author's personal copy
The difference to Fig. 1 in vertical scale used for displaying DIFF12
values should be borne in mind when inspecting Fig. 2, especially for
CO
2
, but the visual impression is nevertheless instructive. The DIFF12
graph for CO
2
always remain positive, between 0.5 and 3 ppm per
year, a consequence of the steady increase in atmospheric CO
2
over
the observation period. The average DIFF12 is about 1.7 ppm/yr,
which as expected adds up to the entire overall increase of 51 ppm
in atmospheric CO
2
between 1980 and 2011. From about 2004, the
average and minimum DIFF12 values for atmospheric CO
2
are some-
what higher than before, but the data series is too short to evaluate
if this is a signicant development or not.
The monthly development of DIFF12 for the two temperature re-
cords is more complicated, ranging from positive to negative values,
reecting repeated periods of warming and cooling over time win-
dows of 12 months. However, the most conspicuous observation is
that all three DIFF12-graphs display variations in concert with each
other, suggesting internal coupling or, alternatively, the common ef-
fect of another factor, inuencing all three variables.
Fig. 3 shows the result of an identical analysis, but here using glob-
al surface air temperature data from the Goddard Institute for Space
Studies (GISS), USA. The general visual impression is identical to
that conveyed by Fig. 2: all three DIFF12 graphs vary in close concert
with each other.
4. DIFF12 surface temperature analyses
A detailed inspection of Figs. 2 and 3 provides some important
clues to what appears to be an important relation between global
temperature and atmospheric CO
2
. For example, the prominent tem-
perature change peak associated with the well-known 1998 El Niño
event is seen to precede the corresponding CO
2
change peak by al-
most 1 yr, and similar lags of several months are seen for most of
the other DIFF12 CO
2
peaks. There are nine main DIFF12 atmospheric
CO
2
peaks visible in Fig. 2, and Table 1 show these nine peak values
for DIFF12 CO
2
and the corresponding nine DIFF12 values for HadSST2
and HadCRUT3. From this it is seen that DIFF12 CO
2
peak values gen-
erally are associated with negative or near zero DIFF12 values for both
HadSST2 and HadCRUT3.
One additional intriguing detail can be observed in both Figs. 2 and
3. The sea surface temperature change peaks typically occursshortly be-
fore the corresponding global surface air temperature change peak,
suggesting a characteristic sequence of events: 1) change of the global
sea surface temperature, 2) change of global surface air temperature,
and 3) change of global atmospheric CO
2
content. In other words, on
this background global temperature changes appear to be initiated at
the surface of the oceans.
Fig. 4 shows the correlation coefcient calculated for the DIFF12
surface temperature (GISS, HadCRUT3 and HadSST2) and global CO
2
data, to investigate the time lag of atmospheric CO
2
in relation to
the three different surface temperature records. The maximum posi-
tive correlation between CO
2
and temperature is found for CO
2
lag-
ging about 9.5 months after GISS, 10 months after HadCRUT3 and
11 months after HadSST2.
In other words, a change of atmospheric CO
2
typically follows about
11 months after the corresponding change of sea surface temperatures,
and 9.510 months later than the global surface air temperature, again
showing that changes in sea surface temperature typically occur a little
(11.5 months) before corresponding changes in the global surface air
temperature. The strongest positive correlation coefcient (0.45) be-
tween CO
2
and temperature is found towards the HadSST2 sea surface
temperature. However, thedifference to surface air temperatures is rel-
atively small, and the maximum positive correlation coefcient is 0.40
for HadCRUT3 and 0.43 for GISS.
5. DIFF12 land and ocean temperature analysis
The NCDC data series on surface air temperatures makes it possi-
ble to analyze the phase relation between atmospheric CO
2
and sub-
sets of surface air temperatures calculated for land, ocean, and
global surface areas, respectively. To achieve this, all three NCDC
datasets were exposed to the DIFF12 numerical procedure described
above. The result is shown in Fig. 5.
The DIFF12 NCDC data for land areas show more marked varia-
tions than the corresponding ocean and global surface air tempera-
ture records. As expected, the ocean record is showing the smallest
temperature change rates, reecting the large difference in heat ca-
pacity between land and ocean surfaces. However, all three DIFF12
NCDC series are seen to vary in concert with each other, as well as
with the DIFF12 CO
2
series. As was the case for the previously inves-
tigated temperature series, changes in the global atmospheric CO
2
data series are seen to lag behind corresponding changes in the
NCDC temperatures.
Fig. 3. 12-month change of global atmospheric CO
2
concentration (NOAA; green), global sea surface temperature (HadSST2; blue) and global surface air temperature (GISS; red
dotted). All graphs are showing monthly values of DIFF12, the difference between the average of the last 12 months and the average for the previous 12 months for each data series.
(For interpretation of the references to color in this gure legend, the reader is referred to the web version of the article.)
54 O. Humlum et al. / Global and Planetary Change 100 (2013) 5169
Author's personal copy
Table 1
Summary of peak DIFF12 CO
2
values and associated DIFF12 HadSST2 and HadCRUT3 values (Fig. 2). Underlining indicates negative values.
Peak
DIFF12 CO
2
123456789
Value DIFF12 CO
2
1.984 2.570 1.472 2.037 2.958 2.552 2.322 2.201 2.383
Value DIFF12 HadSST2 0.004 0.005 0.038 0.059 0.107 0.040 0.063 0.123 0.127
Value DIFF12 HadCRUT3 0.002 0.046 0.030 0.098 0.072 0.005 0.043 0.150 0.139
Fig. 4. Correlation coefcients between DIFF12 monthly surface air temperatures (a: GISS; b: HadCRUT3), sea surface temperature (c: HadSST2) and global atmospheric CO
2
, for
different monthly lags of CO
2
. The maximum positive correlation is found for CO
2
lagging 9.5 months behind GISS, 10 months behind HadCRUT3 and 11 months behind HadSST2.
Numbers in parentheses show the maximum positive correlation coefcient and the associated time lag of CO
2
in months. The grey vertical arrows indicate no lag.
55O. Humlum et al. / Global and Planetary Change 100 (2013) 5169
Author's personal copy
Fig. 6 shows statistics of the CO
2
lag in relation to the three differ-
ent NCDC surface air temperature series, when seen over the entire
study period. The maximum positive correlation (0.45) between tem-
perature and CO
2
is found for changes in atmospheric CO
2
lagging
12 months behind ocean surface air temperature. The lag is shorter
(9.5 months) in relation to land surface air temperature, with a max-
imum positive correlation coefcient of only 0.29. In relation to the
global (land and ocean) NCDC temperature atmospheric CO
2
has a
maximum positive correlation (0.39) for a lag of 10 months after cor-
responding changes in temperature.
Thus, summing up for the analysis of the NCDC data, changes in
atmospheric CO
2
is lagging 9.512 months behind changes in surface
air temperatures calculated for the two main types of planetary sur-
face, land and ocean, respectively. The strongest correlation (0.45)
between atmospheric CO
2
and NCDC temperature is found in rela-
tion to ocean surface air temperatures, suggesting a rather strong
coupling from changes in ocean temperature to changes in atmo-
spheric CO
2
.
6. DIFF12 troposphere temperature analyses
The above observations all suggest the ocean surface temperature
to be an initiating factor in a planetary sequence of events, coupling
temperature and atmospheric CO
2
. Therefore, maintaining a special
focus on the oceans (covering about 71% of the planetary surface),
we proceed to investigate the lag between the changes of tempera-
tures in the lower troposphere above oceans only, compared to asso-
ciated changes in global atmospheric CO
2
. This is shown in Fig. 7,
using A) data on global atmospheric CO
2
and B) data on lower tropo-
sphere temperatures above oceans (UAH). Here tropical oceans are
dened as oceans within 20 degrees north and south of the Equator.
To ensure that our analysis is comparable with the previous analysis,
all four data series were exposed to the DIFF12 yearly change proce-
dure described above.
In general, the DIFF12 temperature change peaks for these three
important ocean regions are seen to be more or less simultaneous
(Fig. 7), but with the lower troposphere temperature change above
the Northern Oceans lagging slightly behind corresponding changes
above the Southern and the Tropical Oceans. This suggests that
ocean surface temperature changes generally are initiated in the
Southern Hemisphere or near the Equator. However, notwithstanding
these comparatively small internal lags in the ocean temperature
changes as reected in the lower troposphere, the corresponding
peaks and lows in the DIFF12 change rate of atmospheric CO
2
are
clearly lagging several months after the lower troposphere tempera-
ture change above all three main ocean regions.
Fig. 8 shows the correlation coefcient between DIFF12 for global
atmospheric CO
2
and DIFF12 for lower troposphere temperature
(UAH) over land, ocean and global, respectively, for different lags of
CO
2
in relation to temperature. The time lag is 8 months for land
areas, 10 months for oceans, and 9 months for both land and oceans.
The time lag of CO
2
is slightly shorter (1 month) than found for the
surface temperature records, but changes in atmospheric CO
2
clearly
lag the associated changes in all three lower troposphere tempera-
tures analyzed. The correlation coefcient for the global 9-month
lag is 0.48.
The overall shorter time lag of CO
2
in relation to lower tropo-
sphere temperatures, compared to surface temperatures, suggests a
typical sequence of temperature change events starting at the planet
surface and propagating to the lower troposphere with about one
month delay, given the time resolution of the data series.
7. DIFF12 hemispheric temperature analyses
The above observations suggest the ocean surface temperature to
be an initiating factor in a planetary sequence of events, coupling
temperature and atmospheric CO
2
. This motivates an analysis of the
relation between global atmospheric CO
2
and surface air temperature
in the two hemispheres separately (Fig. 9), as especially the surface of
the Southern Hemisphere is dominated by oceans, and the lower tro-
posphere temperature change above the northern oceans appears to
be lagging slightly behind corresponding changes above the southern
and the tropical oceans.
Here we show hemispherical temperatures according to NCDC, but
other temperature records (HadCRUT3 and GISS) show essential an
identical picture. The visual impression gained from Fig. 9 is similar to
what is shown by the analysis above: DIFF12 variations for atmospheric
CO
2
are tracking behind both sets of hemispherical surface air tempera-
tures. However, because of the dominance of oceans in the Southern
Hemisphere, the temperature change rates are smaller here than in the
Northern Hemisphere. An inspection of Fig. 9 reveals that each DIFF12
temperature peak for the Southern Hemisphere is followed by a corre-
sponding DIFF12 peak in global atmospheric CO
2
. Usually Northern
Hemisphere DIFF12 peaks occur in concert with Southern Hemisphere
peaks, but there are a few occasions where a Northern Hemisphere
DIFF12 temperature peak occurs independently (e.g., early 1986), and
Fig. 5. 12-month change of global atmospheric CO
2
concentration (NOAA; green), change in global surface air temperature (NCDC; blue), land surface air temperature (NCDC; yel-
low) and ocean surface air temperature (NCDC; red dotted). All graphs are showing monthly values of DIFF12, the difference between the average of the last 12 months and the
average for the previous 12 months for each data series. (For interpretation of the references to color in this gure legend, the reader is referred to the web version of the article.)
56 O. Humlum et al. / Global and Planetary Change 100 (2013) 5169
Author's personal copy
is not followed by a corresponding DIFF12 CO
2
peak. By this the visual
analysis thus favors a mainly Southern Hemisphere coupling to DIFF12
CO
2
peaks.
Fig. 10 shows the correlation coefcient between DIFF12 for global
atmospheric CO
2
and DIFF12 for Northern and Southern Hemisphere
temperature (NCDC), respectively, for different lags of CO
2
in relation
to temperature. Positive correlation is found only for DIFF12 CO
2
lag-
ging DIFF12 temperature. The maximum positive correlation is
found for DIFF12 atmospheric CO
2
lagging about 9.5 months behind
Northern Hemisphere temperature, and 11 months behind Southern
Hemisphere temperature, suggesting the Southern Hemisphere to be
leading the dynamics shown in Fig. 9. The correlation coefcient is
considerably higher (0.56) for the Southern Hemisphere than for
the Northern Hemisphere (0.26), indicating the association between
changes in hemispherical temperature and changes i n global atmospheric
CO
2
to be especially strong for the Southern Hemisphere. Thus, both anal-
yses suggest a mainly Southern Hemisphere origin of observed DIFF12
changes for atmospheric CO
2
.
Fig. 6. Correlation coefcients between DIFF12 monthly NCDC surface air temperatures (a: land; b: ocean; c: global) and global atmospheric CO
2
, for different monthly lags of CO
2
.
The maximum positive correlation is found for CO
2
lagging 9.5 months behind land surface air temperature, 12 months behind ocean surface air temperature and 10 months be-
hind the global surface air temperature. Numbers in parentheses show the maximum positive correlation coefcient and the associated time lag of CO
2
in months. The grey vertical
arrows indicate no lag.
57O. Humlum et al. / Global and Planetary Change 100 (2013) 5169
Author's personal copy
8. DIFF12 release of CO
2
from anthropogene sources
We next investigated the degree of correlation between atmo-
spheric CO
2
and the release of CO
2
from different anthropogene
sources, using monthly data from the Carbon Dioxide Information
Analysis Center (CDIAC). DIFF12 values were calculated as above,
and the result is shown in Fig. 11. Since 2009 only annual data are
at hand, and these were therefore extrapolated on a at average
monthly basis. The inclusion of these data in the present analysis is
important, as the increase in atmospheric CO
2
since 1958 usually is
ascribed to anthropogene release of CO
2
from anthropogene sources.
It is generally assumed that about half of the emitted anthropogene
CO
2
is absorbed by the oceans and on land by various physical and
chemical processes, leaving about the remaining half to explain the
rise in global atmospheric CO
2
.
Fig. 11 shows visually that the coupling between DIFF12 changes
in anthropogene CO
2
and DIFF12 changes in atmospheric CO
2
is
weak, especially when compared to what was the case for tempera-
tures (Figs. 2, 3, 5, 7 and 9). Several of the change rate peaks in
anthropogene CO
2
appear to lag behind peaks in atmospheric CO
2
.
In other cases, the direction of change is opposed to each other, as
is exemplied by the pronounced 1998 El Niño peak in atmospheric
CO
2
, corresponding to almost zero change in anthropogene CO
2
. The
2011 change peak for atmospheric CO
2
is another example, as this oc-
curs simultaneously with the negative change rate for anthropogene
CO
2
, caused by the recent economical crisis. A similar contrast has
previously been pointed out (Jaworowski et al., 1992) for Mauna
Loa CO
2
data during the petroleum crisis 197374.
As anthropogene CO
2
generally is assumed to explain the modern
rise in atmospheric CO
2
, DIFF12 for this is expected to lag somewhat
after DIFF12 for anthropogene CO
2
. In contrast to this, Fig. 11 appears
to suggest the opposite relation, if anything. This might partly be due
to the extraordinary high use of fossil fuels especially in the Northern
Hemisphere for heating purposes in relatively cold periods, which as-
sociates with DIFF12 CO
2
lows (Fig. 2). However, a detailed analysis of
this falls beyond the present investigation.
In general the coupling appears visually weak compared to the pre-
vious comparisons with temperature (Figs.2,3,5,7and9). The relation
between the two CO
2
data series also appears partly conicting, and is
difcult to determine precisely on a visual basis alone. However, it is ev-
ident from the visual analysis that changes in atmospheric CO
2
are gen-
erally not tracking changes in anthropogene emissions, which is
contrary to expectation, if anthropogene CO
2
is the main driver for the
observed rise in global atmospheric CO
2
.
The release of CO
2
from fossil fuels into the atmosphere is
strongly asymmetrical, with the overwhelming majority being re-
leased in the Northern Hemisphere around 40°N (Fig. 12). This mo-
tivates a hemispherical analysis of the relation between changes in
anthropogene CO
2
released and changes in atmospheric CO
2
.Due
to the asymmetrical release pattern the inuence of anthropogene
CO
2
would be expected to be stronger and more direct in the North-
ern Hemisphere, compared to the Southern Hemisphere.
Fig. 13 shows the relation between DIFF12 for anthropogene CO
2
and DIFF12 values for atmospheric CO
2
recorded at four different sta-
tions representing a transect from the High Arctic to the South Pole.
Here we are using the period from July 1991 to December 2006, as
data (NOAA) only are available from all four stations for this time
window. Once again, there is no direct visual coupling from the
DIFF12 values for anthropogene CO
2
to any of the four individual re-
cords, but instead the DIFF12 values for anthropogene CO
2
appears
to lag somewhat after DIFF12 values for all individual records.
Comparing the dynamic behavior of the four DIFF12 series for at-
mospheric CO
2
, a systematic sequence however stands out: peaks
and lows in DIFF12 values for Ascension Island (8°S) tend to occur
shortly before the DIFF12 values for Mauna Loa (20°N), which leads
before the DIFF12 values for the South Pole (90°S), which leads before
the DIFF12 values for Alert (82°N) in the High Arctic. From this,
changes in atmospheric CO
2
appear to be initiated near or a short dis-
tance south of the Equator, and from there spread towards the two
poles within a year or so. En route, the signal presumable is modulat-
ed by local and regional effects, as is indicated by the much larger an-
nual CO
2
variation (not shown here) in the High Arctic, compared to
that recorded at the South Pole. There is however no indications of
the main signal originating at mid-latitudes in the Northern Hemi-
sphere as would be expected from the release pattern shown in
Fig. 12. The above visual analysis might be taken further by subse-
quent analyses, but here is sufces to investigate if correlation analy-
ses between the four DIFF12 atmospheric CO
2
records and the DIFF12
anthropogene CO
2
record support the result of the visual analysis.
Fig. 14 shows the calculated correlation coefcients between
changes in anthropogene CO
2
and changes in atmospheric CO
2
for the
four stations shown in Fig. 13. In all four cases there is a negative corre-
lation from the time of release and 1724 months later between DIFF12
changes in anthropogene CO
2
and DIFF12 changes in atmospheric CO
2
,
Fig. 7. 12-month change of global atmospheric CO
2
concentration (NOAA; green), change in lower troposphere temperature over oceans in the Northern Hemisphere (UAH; blue),
in the Southern Hemisphere (UAH; yellow) and in the Tropics (UAH; red dotted). All graphs are showing monthly values of DIFF12, the difference between the average of the last
12 months and the average for the previous 12 months for each data series. (For interpretation of the references to color in this gure legend, the reader is referred to the web
version of the article.)
58 O. Humlum et al. / Global and Planetary Change 100 (2013) 5169
Author's personal copy
showing that changes in the emission of anthropogene CO
2
are not
causing changes in atmospheric CO
2
. The strongest correlation coef-
cient ranges from 0.30 to 0.52, and the corresponding lag from 4
to 19 months, depending on latitude. The maximum absolute value of
the negative correlation coefcient is larger for the Southern Hemi-
sphere than for the Northern Hemisphere. On the other hand, the lag
of DIFF12 for anthropogene CO
2
in relation to DIFF12 for atmospheric
CO
2
is larger in the Northern Hemisphere than in the southern. Thus,
there is no indication of a northern mid-latitude origin for DIFF12
changes in atmospheric CO
2
. The lag is about 4 months for Ascension
(8°S), 9 months for the South Pole (90°S), 14 months for Mauna Loa
(20°N), and about 19 months for Alert (82°N) in the High Arctic.
As each station is correlated with identical values (DIFF12 for
anthropogene CO
2
) the observed lags suggest a sequence of events
starting near Equator in the Southern Hemisphere, and from there
propagating towards the two poles.
Fig. 15 (panel a) shows the calculated correlation coefcient
between changes in anthropogene CO
2
and changes in global
Fig. 8. Correlation coefcients between DIFF12 monthly lower troposphere temperatures (UAH) and global atmospheric CO
2
, for different monthly lags of CO
2
in relation to tem-
perature. The maximum positive correlation is found for CO
2
lagging 9 months after lower global troposphere temperatures, 10 months after lower troposphere temperatures over
oceans, and 8 months after lower troposphere temperatures over land. Numbers in parentheses show the maximum positive correlation coefcient and the associated time lag of
CO
2
in months. The grey vertical arrows indicate no lag.
59O. Humlum et al. / Global and Planetary Change 100 (2013) 5169
Author's personal copy
atmospheric CO
2
for the entire period 19802011. It is seen that the
correlation generally is low as was suggested by the visual analysis
above, and noticeably smaller than found for correlations with tem-
peratures (Figs. 4, 6, 8 and 10). From the time of release and to
about 1 yr later there is a negative correlation between DIFF12
changes in anthropogene CO
2
and DIFF12 changes in atmospheric
CO
2
. The maximum positive correlation is found by correlating
DIFF12 changes in CO
2
from fossil fuels with atmospheric CO
2
Fig. 9. 12-monthchange of global atmosphericCO
2
concentration (NOAA; green), change in Northern Hemisphere surface air temperature(NCDC; red dotted),and Southern Hemisphere
air temperature (NCDC; blue).All graphs are showingmonthly values of DIFF12, the differencebetween the average of thelast 12 months andthe average for the previous 12 monthsfor
each data series. (For interpretation of the references to color in this gure legend, the reader is referred to the web version of the article.)
Fig. 10. Correlation coefcientsbetween DIFF12 monthlysurface air temperatures in the northern (a) and the SouthernHemisphere (b)and global atmosphericCO
2
, for differen t monthly
lags of CO
2
. The maximum positive correlationis found for CO
2
lagging9.5 months behindNorthern Hemispheretemperature, and 11 months behind SouthernHemisphere temperature.
Numbers in parentheses show the maximum positive correlation coefcient and the associated time lag of CO
2
in months. The grey vertical arrows indicate no lag.
60 O. Humlum et al. / Global and Planetary Change 100 (2013) 5169
Author's personal copy
DIFF12 changes 11 months earlier. However, this is opposite to what
would be expected if changes in anthropogene CO
2
had signicant
impact on changes in the amount of atmospheric CO
2
.
However, on shorter time scales the relation is highly variable as
the two panels b and c in Fig. 13 show. Dividing the entire observation
period into the two sub-periods 19801998 and 19982011 yields a
different result as to the correlation between DIFF12 changes in
anthropogene CO
2
and DIFF12 changes in global atmospheric CO
2
(panel a). For the period 19801998 the correlation is found to be
positive for the period at and following release, while it becomes neg-
ative for the time window 19982011. A positive correlation would
indeed be expected, if changes in the release of anthropogene CO
2
were controlling changes in atmospheric CO
2
. It is however contrary
to expectation that the maximum positive correlation is found
1 month before the time of release (Fig. 15, panel b).
This demonstrates that the relation between changes in the re-
lease of CO
2
from anthropogene sources and changes in atmospheric
CO
2
is not stable, but undergoes signicant changes on a decadal
time scale. Had changes in the release of anthropogene CO
2
repre-
sented the main control on changes in atmospheric CO
2
,therelation
would be stable. A possible explanation for the unstable correlation
might be due to the combined El NiñoLa Nino event 19972001,
affecting the two investigated sub periods (Fig. 15) differently. An-
other possible explanation was provided by Njau (2007), suggesting
that anthropogene CO
2
has so far been added to the atmosphere in an
amount which would have otherwise been added naturally, espe-
cially by the warming oceans, had the emission of anthropogene
CO
2
been absent. However, a detailed analysis of such and other pos-
sibilities falls beyond the present investigation, and here it sufces to
conclude that the absence of a stable, positive correlation suggests
other processes than the emission of anthropogene CO
2
to control
the main features of the observed changes in atmospheric CO
2
.
Summing up,our analysis suggests that changes in atmospheric CO
2
appear to occur largely independently of changes in anthropogene
emissions. A similar conclusion was reached by Bacastow (1976),
suggesting a coupling between atmospheric CO
2
and the Southern
Oscillation. However, by this we have not demonstrated that CO
2
re-
leased by burning fossil fuels is without inuence on the amount of
atmospheric CO
2
, but merely that the effect is small compared to
the effect of other processes. Our previous analyses suggest that
such other more important effects are related to temperature, and
with ocean surface temperature near or south of the Equator
pointing itself out as being of special importance for changes in the
global amount of atmospheric CO
2
.
Fig. 11. 12-month change of global atmospheric CO
2
concentration (NOAA, green), and the change in release of CO
2
from burning of fossil fuels (CDIAC, red). Both graphs are show-
ing monthly values of DIFF12, the difference between the average of the last 12 month and the average for the previous 12 months for each data series.
Fig. 12. Latitudinal 1990 summary of CO
2
emissions from anthropogene sources. Data source: Carbon Dioxide Information Analysis Center (CDIAC).
61O. Humlum et al. / Global and Planetary Change 100 (2013) 5169
Author's personal copy
9. Volcanic eruptions
Volcanic eruptions are obviously relevant also to consider here, as
they may contribute not only with tephra (e.g. ash), but also with
substantial amounts of greenhouse gases such as H
2
O and CO
2
to
the atmosphere, inuencing both temperature and the total amount
of greenhouse gasses. In addition, a gas like SO
2
is injected into the at-
mosphere during eruptions. Oxidation and gas-to-particle conversion
transforms SO
2
to sulphuric acid (H
2
SO
4
), which may effect cloud for-
mation. Each year several volcanic eruptions occur globally, but the
majority are small, and not signicant within the present global
context.
The reported magnitude of volcanic eruptions depends much on
both the experience and vantage point of the observer. To meet the
need for a meaningful magnitude measure that can be easily applied
to eruption sizes, Newhall and Self (1982) therefore integrated quan-
titative data with the subjective descriptions of observers, resulting in
the Volcanic Explosivity Index (VEI) to provide a relative measure of
the explosiveness of volcanic eruptions. The VEI scale is a simple
08 index of increasing explosivity, based on volume of ejected
mass, eruption cloud height, and other qualitative observations. The
scale is logarithmic, with each interval on the scale representing a
tenfold increase in observed ejecta criteria, with the exception of be-
tween VEI 0, VEI 1 and VEI 2. The VEI-divisions indicate the amount of
tephra (volcanic ash) ejected into the atmosphere, but do not inform
directly about the amount of water vapour and CO
2
emitted, but usu-
ally the amount of these is expected to follow roughly the VEI class,
although with individual variations (see below). The most important
characteristics of the individual VEI-classes are shown in Table 2.
Here we consider only eruptions greater than VEI size class 4, as
these eruptions all are classied as large and therefore may have
ejected signicant of tephra and other eruption products to high alti-
tudes, across the tropopause and into the stratosphere. The timing of
such major volcanic eruptions since 1981 is shown in Fig. 16, together
with the DIFF12 changes in the amount of atmospheric CO
2
. Data on
volcanic eruptions were obtained from the Global Volcanism Program
(GWP).
As the VEI-data are qualitatively different from all previous data
types used in the present analysis, being somewhat subjective and
representing discrete events, we shun from attempting any kind of
statistical analysis of these data, and will only consider briey the vi-
sual impression gained from Fig. 16.
It is well known that major volcanic eruptions may inuence glob-
al temperatures 45 yr (see, e.g., Lockwood and Fröhlich, 2008;
Thompson et al., 2009) after the culmination of the eruption, al-
though with variations reecting the position of the volcano in rela-
tion to the Equator. The variable temperature effect of volcanic
eruptions is illustrated by comparing the timing of major eruptions
shown in Fig. 16 with the temperature graphs shown in Figs. 1, 2, 3,
5 and 7. For example, these diagrams all show global cooling after
the very large Mount Pinatubo eruption that began in earnest June
1991. However, Fig. 16 shows that this very large eruption not to
have resulted in any increase of atmospheric CO
2
. The main DIFF12
change observed is actually a period of low growth rate for atmo-
spheric CO
2
. The same observation applies for the group of 2008 erup-
tions. This effect on CO
2
may be explained by volcanic gasses/aerosols
and airborne debris affecting cloud cover, leading to cooling of the
uppermost part of the oceans and therefore increased ability to take
up atmospheric CO
2
, according to Henry's law.
Summing up, although there has been several volcanic eruptions
of at least VEI 4 during the study period, the direct effect on atmo-
spheric CO
2
in general appears to be limited on the time scale inves-
tigated, and only contributes little to explain directly observed
changes in atmospheric CO
2
within the study period. Perhaps erup-
tions larger th an VEI = 6 are r equired t o affe ct the global atmospher-
ic composition and climate signicantly, but there has been no
eruptions of this magnitude during the study period. The analysis
might be taken further by also considering the type of volcanism,
Fig. 13. 12-month change of global atmospheric CO
2
concentration for Alert (NOAA; blue), Mauna Loa (NOAA; green), Ascension Island (NOAA; yellow), and the south Pole (NOAA;
purple), and the change in release of anthropogene CO
2
(CDIAC; red). All graphs show monthly values of DIFF12, the difference between the average of the last 12 months and the
average for the previous 12 months for each data series. (For interpretation of the references to color in this gure legend, the reader is referred to the web version of the article.)
Fig. 14. Correlation coefcients between DIFF12 change in release of anthropogene CO
2
(CDIAC) and DIFF12 change in atmospheric CO
2
at different stations, for different lags of CO
2
in
relation to CO
2
from fossil fuel. The maximum negative correlation is found for changes in atmospheric CO
2
ranges from 4 to 19 months after changes in CO
2
release from fossil fuels.
Numbers in parentheses show the maximum absolute correlation coefcient and the associated time lag ofCO
2
in months. The grey vertical arrows indicate no lag (the time of release).
62 O. Humlum et al. / Global and Planetary Change 100 (2013) 5169
Author's personal copy
63O. Humlum et al. / Global and Planetary Change 100 (2013) 5169
Author's personal copy
and from which depth in the crust it originates. In general,
deep-seated volcanism is relatively rich in CO
2
, while tholeiitic erup-
tions originating from more shallow depths usually are relatively
rich in H
2
O.
As was the case for anthropogene CO
2
, we can therefore not con-
clude that volcanic activity since January 1980 is without inuence on
atmospheric CO
2
, but only that the effect of eruptions in the study peri-
od apparently is small compared to other factors. On the time scale in-
vestigated here, the dominant effect of volcanic eruptions appears to
be increased removal of CO
2
from the atmosphere by the oceans, pre-
sumably caused by volcanic eruptions affecting the global cloud cover,
and thereby resulting in lower ocean surface temperature. On longer
(geological) time scales the relation may however well be different.
10. Fourier frequency analyzes
Visual inspection of Figs. 2, 3, 5, 7 and 9 suggests that some DIFF12
series may be changing according to one or several rhythmic varia-
tions. Therefore, to investigate if these records are in fact character-
ized by such periodic variations, a Fourier frequency analysis was
carried out on four key data series, 1) atmospheric CO
2
(NOAA), 2)
Fig. 15. Correlation coefcients between DIFF12 change in release of anthropogene CO
2
(CDIAC) and global atmospheric CO
2
, for different lags of atmospheric CO
2
in relation to CO
2
from fossil fuel. Numbers in parentheses show the maximum positive correlation coefcient and the associated time lag of CO
2
in months. The grey vertical arrows indicate no lag
(the time of release).
64 O. Humlum et al. / Global and Planetary Change 100 (2013) 5169
Author's personal copy
HadCRUT3, 3) HadSST2 and 4) CO
2
released by use of fossil fuels
(CDIAC). The results are shown in Fig. 17.
The DIFF12 atmospheric CO
2
record is seen to be dominated by
two periods of about 3.8 and 2.5 yr length, both with amplitude
greater than 0.25 ppm. The two DIFF12 temperature spectra are rather
similar to each other, dominated by a period of about 3.7 yr length with
amplitude greater than 0.09 °C, and both showing another important
peak around 2.4 yr. The 2.52.4 yr period seen in the three uppermost
records in Fig. 17 may possibly reect the inuence of the Southern
Oscillation, as suggested by both Lamb (1972);Bacastow (1976).
However, the spectrum for the DIFF12 CO
2
anthropogene record
(lowermost panel in Fig. 13) is different. In particular, there is no fre-
quency peak suggesting the existence of an important period in the
range from 3.7 to 3.8 yr, which is a principal period length in the
other three records, but the anthropogene CO
2
record does show a
less pronounced peak around 4 yr. The potential implication of this
peak will be commented later.
The signicance levels shown in Fig. 17 are peak-based critical
limit signicance levels, which are of particular merit in ascertaining
the signicance of the largest spectral component (SeaSolve, 2003;
Humlum et al., 2012). In this type of test, one seeks to disprove the
null hypothesis postulating either a white noise signal with no auto-
correlation (AR(1)=0.0), or a red noise signal with autocorrelation
(AR(1)>0.0). Red noise is present when the background power de-
creases with increasing frequency, and the autoregressive value is a
measure of the similarity between observations in a time series as a
function of the time separation between them. All four DIFF12 series
are characterized by high autocorrelation, with AR(1) coefcients
ranging from 0.97 to 0.99.
With the background set, the peak spectral power was then com-
pared against the various critical limits. A 99% critical limit is that
level where in only 1 of 100 separate random noise signals the
highest peak would achieve this height strictly due to random chance.
Likewise, for a peak reaching the 50% critical limit there is a 5050
probability that this could have arisen strictly from chance. The criti-
cal limit test used here is different from traditional condence or sig-
nicance levels that apply to a single data set only. For example, a
standard 95% condence limit would specify a level where 5% of the
points in a single spectrum would be expected to lie above this height
strictly due to random chance.
Exposed to the critical limit test all four DIFF12 series display sev-
eral frequency peaks exceeding the 95% critical limit. For that reason
the alternative null hypothesis, that the observed oscillations repre-
sent red noise only, can be rejected using a 95% peak-based critical
limit test, and the signicant frequency peaks are positively worthy
of interpretation. Several peaks are seen to exceed even the 99% crit-
ical limit.
Comparing the different Fourier frequency spectra shown in
Fig. 17, the degree of similarity between the two temperature records
(HadCRUT3; HadSST2) and the atmospheric CO
2
record appears
higher than the similarity between anthropogene CO
2
and atmo-
spheric CO
2
. This lends support to the result of the previous analyzes,
that changes in atmospheric CO
2
appear to be coupled especially to
changes in temperature, and not to the same degree to changes in
CO
2
released by anthropogene sources.
The lower panel (d) in Fig. 17 indirectly shows the main rhythms of
the world economy. However, theamount of anthropogene CO
2
must to
some degree also reect the production of heat and electricity for
heating purposes, especially during winter in the Northern Hemisphere,
as this is where most people live. Thus, it is not entirely surprising that
the prominent temperature DIFF12 peak of about 3.7 yr (HadCRUT3
and HadSST2) may nd itself represented in the broad frequency peak
around 0.25 yr
1
(4 yr) in the DIFF12 data for anthropogene CO
2
.The
visual inspection of Fig. 11 leads to a similar interpretation.
Table 2
Volcanic Explosivity Index (VEI) eruption size classes based on Newhall and Self (1982) and Simkin and Siebert (1994).
VEI class 0 1 2 3 4 5 6 7 8
Description Non-explosive Small Moderate Moderatelarge Large Very large Very large Very large Very large
Tephra (m
3
)b1×10
4
b1×10
6
b1×10
7
b1×10
8
b1×10
9
b1×10
10
b1×10
11
b1×10
12
>1×10
12
Cloud height (km) b0.1 0.1115315 1025 >25 > 25 >25 > 25
Fig. 16. 12-month change of global atmospheric CO
2
concentration (NOAA; green), and the timing of major (VEI > 3) volcanic eruptions (GWP). The CO
2
graph is showing monthly
values of DIFF12, the difference between the average of the last 12 months and the average for the previous 12 months for each data series. The DIFF12 CO
2
values have been shifted
12 months forward to correspond correctly with the time axis used for plotting the volcanic eruptions.
65O. Humlum et al. / Global and Planetary Change 100 (2013) 5169
Author's personal copy
11. Discussion
By the above analysis we have demonstrated that there exists a
clear phase relationship between the change of different global
temperature indices, and corresponding changes in the global
amount of atmospheric CO
2
. We have used data on CO
2
and temper-
ature with a monthly time resolution for the time window January
1980December 2011, but instead of analyzing the monthly data,
Fig. 17. Fourier frequency spectra (Best Exact N composite algorithm; SeaSolve, 2003) for DIFF12's for changes in a) atmospheric CO
2
, b) HadCRUT3, c) HadSST2 and d) CO
2
released
by use of fossil fuels. The DIFF12 atmospheric CO
2
record is dominated by highly signicant periods of about 3.8 and 2.5 yr length, while especially a period of about 3.7 yr is dom-
inant in both DIFF12 temperature records. The DIFF12 record for CO
2
from fossil fuels is characterized by a different type of spectrum, and with peaks of generally less signicance
than characterizing the other records. The grey tone indicates increasing amplitude within each record. The stippled lines indicate peak-based critical limit signicance levels.
66 O. Humlum et al. / Global and Planetary Change 100 (2013) 5169
Author's personal copy
we have removed the inuence of annual cycles by analyzing DIFF12
values. The average DIFF12 CO
2
value for the entire observation peri-
od is about 1.7 ppm/yr, which adds up to the entire observed increase
of 51 ppm of atmospheric CO
2
since January 1980. Therefore, to the
degree that the observed DIFF12 atmospheric CO
2
values (Figs. 3, 5,
7 and 9) can be explained by temperature changes; the overall in-
crease of atmospheric CO
2
since January 1980 (Fig. 1) is also
explained.
The coupling from changes in atmospheric CO
2
to changes in tem-
perature appears to be weak. The rate of temperature increase often
decreases when the amount of atmospheric CO
2
increases (Figs. 2,
3, 5, 7 and 9). There are even several examples where the DIFF12 tem-
perature change becomes negative (Table 1), corresponding to de-
creasing temperatures, even though the contemporary DIFF12 CO
2
change is positive.
In general, we nd that changes in atmospheric CO
2
are lagging
behind changes in any of the ve different temperature records con-
sidered. The typical lag is 9.512 months for surface temperatures
and about 9 months for lower troposphere temperatures, suggesting
a temperature sequence of events from the surface to the lower
troposphere.
As cause always must precede effect, this observation demon-
strates that modern changes in temperatures are generally not in-
duced by changes in atmospheric CO
2
. Indeed, the sequence of
events is seen to be the opposite: temperature changes are taking
place before the corresponding CO
2
changes occur.
As the theoretical initial temperature effect of changes in atmo-
spheric CO
2
must materialize rst in the troposphere, and then subse-
quently at the planet surface (land and ocean), our diagrams 28
reveal that the common notion of globally dominant temperature
controls exercised by atmospheric CO
2
is in need of reassessment.
Empirical observations indicate that changes in temperature general-
ly are driving changes in atmospheric CO
2
, and not the other way
around.
Numerical global climate models generally assume atmospheric
CO
2
in combination with alleged feed-back effects on atmospheric
humidity and cloud cover to have a clear net warming effect, and
that changes in atmospheric CO
2
therefore represent a main driver
for global temperature changes. For that reason changes in tempera-
ture should therefore be lagging behind corresponding changes in
CO
2
. However, Figs. 4, 6, 8 and 10 show correlation between changes
in temperature and CO
2
to be negative for negative offsets (tempera-
ture lagging CO
2
), indicating that changes towards higher concentra-
tions of atmospheric CO
2
then empirically would associate with less
rapid temperature increase or even a temperature decrease. Howev-
er, as this would invalidate the basic assumption of CO
2
having a
clear net warming effect, the perception of temperature lagging be-
hind CO
2
must therefore be rejected. A visual inspection of the data
displayed in Figs. 2, 3, 5, 7 and 9 also show the notion of temperature
lagging CO
2
to be implausible.
Thus, the simplest explanation of observed changes in DIFF12 for at-
mospheric CO
2
is that they are induced by changes in temperature, as
illustrated by Figs. 210. Consequently, a substantial part of the atmo-
spheric increase of CO
2
since January 1980 can be explained by associat-
ed changes in temperature,and presumably especially changes inocean
temperatures (Toggweiler, 1999; Monnin et al., 2001; Goldberg, 2008),
as this is where we nd both the strongest correlation to changes in CO
2
(Figs. 4, 6 and 8), and the longest time lag.
The maximum positive correlation between ocean temperatures
and atmospheric CO
2
is within a range from 0.45 to 0.48, depending
on which dataset is considered (HadSST2, NCDC or UAH). With a sam-
ple size of 361 (number of monthly DIFF12 values) these correlation
coefcients are highly signicant at the 0.05 level, and correspond
to a goodness-of-t(r
2
) ranging from 0.20 to 0.23. This represents a
fair degree of explanation, and far bigger than achieved by any
other factor considered in the present analysis, but it also suggests
that there are other factors beyond ocean surface temperature
which have inuenced observed changes in atmospheric CO
2
since
January 1980. Examples of such potential factors are changes in soil
moisture, living biomass, volcanic eruptions, geological weathering
processes, burning of fossil fuels, etc. The correlation between CO
2
re-
leased by anthropogene sources and changes in atmospheric CO
2
is not
stable (Fig. 15), and not able toexplain much of the observed increase in
atmospheric CO
2
since January 1980. A qualitatively identical conclusion
may possibly be suggested for the effect of volcanic eruptions during the
study period, but the character of the volcanic data available does not
make it possible to carry out a comparable statistical analysis on this. Ac-
tually, on the time scale investigated, the net effect of a major volcanic
eruption appears to be a reduction of the prevailing increase rate of at-
mospheric CO
2
, probably an effect of ocean cooling induced from cloud
effects.
Summing up, monthly data since January 1980 on atmospheric CO
2
and sea and air temperatures unambiguously demonstrate the overall
global temperature change sequence of events to be 1) ocean surface,
2) surface air, 3) lower troposphere, and with changes in atmospheric
CO
2
always lagging behind changes in any of these different tempera-
ture records.
A main control on atmospheric CO
2
appears to be the ocean sur-
face temperature, and it remains a possibility that a signicant part
of the overall increase of atmospheric CO
2
since at least 1958 (start
of Mauna Loa observations) simply reects the gradual warming of
the oceans, as a result of the prolonged period of high solar activity
since 1920 (Solanki et al., 2004). Based on the GISP2 ice core proxy re-
cord from Greenland it has previously been pointed out that the pres-
ent period of warming since 1850 to a high degree may be explained
by a natural c. 1100 yr periodic temperature variation (Humlum et al.,
2011).
The atmospheric CO
2
growth rate DIFF12 is seen to be at maximum
about the same time when SST DIFF12 is on the falling limb from a pre-
vious peak, indicating ocean temperature to be approaching a tempera-
ture maximum or falling after passing a peak (see, e.g., Fig. 2 and
Table 1). As the rate of CO
2
net outgassing from the ocean then is affect-
ed by reduced solubility, this offers a simple physical explanation of the
observed time lag. At least, the association between periods of maxi-
mum DIFF12 CO
2
increase and no or negative ocean surface tempera-
ture change (Table 1)isdifcult to reconcile with the notion of
atmospheric CO
2
changes controlling changes in ocean surface temper-
ature. However, more work is needed to investigate the details of
this, and it is worth noticing that atmospheric CO
2
increases over
the period considered, even when air temperatures decreases. It is
however clear from the data that emission of CO
2
from the oceans
and other natural sources plays an important role in observed
changes of atmospheric CO
2
. Independent of human emissions, this
contribution to atmospheric CO
2
is not controllable and little pre-
dictable, although the coupling to ocean surface temperature dem-
onstrated above may provide a starting basis for future prediction
attempts.
Analyses of a pole-to-pole transect of atmospheric CO
2
records
suggest that changes in atmospheric CO
2
are initiated south of the
Equator, but probably not far from the Equator, and from there
spreads towards the two poles within a year or so (Fig. 13). This ob-
servation specically points towards the oceans at or south of the
Equator as an important source area for observed changes in atmo-
spheric CO
2
. The major release of anthropogene CO
2
is taking place
at mid-latitudes in the Northern Hemisphere (Fig. 12), but the
northsouth transect investigated show no indication of the main
change signal in atmospheric CO
2
originating here. The main signal
must therefore be caused by something else. A similar conclusion,
but based on studies of the residence time of anthropogenic CO
2
in
the atmosphere, was reached by Segalstad (1998);Essenhigh (2009).
In general, the level of atmospheric CO
2
is slightly higher in the
Northern Hemisphere than in the Southern Hemisphere. This might
67O. Humlum et al. / Global and Planetary Change 100 (2013) 5169
Author's personal copy
be seen as due to anthropogene CO
2
mainly being released in the
Northern Hemisphere (Fig. 12). However, had this being the explana-
tion, the northern mid latitudes should had been the origin for
changes in atmospheric CO
2
, which is shown not to be the case
(Figs. 13 and 14). Thus, presumably other processes are responsible
for the northsouth gradient in atmospheric CO
2
, such as, e.g., respi-
ration by land-based organisms (especially Northern Hemisphere)
and ocean uptake (especially Southern Hemisphere).
Over the entire study period atmospheric CO
2
shows a continuous
increase, when annual variations are ignored. This might also be
interpreted as being the result of the release of anthropogene CO
2
,
but the observed propagation of the main atmospheric CO
2
change
signal along the pole-to-pole transect (Fig. 13) seems to argue against
such an interpretation. The signal propagation instead suggests a pos-
sible connection to especially the southern oceans and their surface
temperature, but a detailed analysis of this falls beyond the present
study.
The modern relation between temperature and CO
2
is qualitatively
identical to that demonstrated by ice cores for the Quaternary glacial
interglacial transitions (Mudelsee, 2001; Caillon et al., 2003), although
the modern time lag between temperature and CO
2
is considerably
shorter. However, this is presumably reecting the much coarser time
resolution provided by ice cores, displaying only changes on a
multi-decadal scale. This is partly due to sampling resolution, partly
due to gas diffusion within the ice that averages out any surfacetemper-
ature variability shorter than a few decades (Severinghaus et al., 1998).
12. Conclusions
There exist a clear phase relationship between changes of atmo-
spheric CO
2
and the different global temperature records, whether
representing sea surface temperature, surface air temperature, or lower
troposphere temperature, with changes in the amount of atmospheric
CO
2
always lagging behind corresponding changes in temperature.
(1) The overall global temperature change sequence of events ap-
pears to be from 1) the ocean surface to 2) the land surface to
3) the lower troposphere.
(2) Changes in global atmospheric CO
2
are lagging about 11
12 months behind changes in global sea surface tempera-
ture.
(3) Changes in global atmospheric CO
2
are lagging 9.510 months
behind changes in global air surface temperature.
(4) Changes in global atmospheric CO
2
are lagging about 9 months
behind changes in global lower troposphere temperature.
(5) Changes in ocean temperatures appear to explain a substantial
part of the observed changes in atmospheric CO
2
since January
1980.
(6) CO
2
released from anthropogene sources apparently has little in-
uence on the observed changes in atmospheric CO
2
,and
changes in atmospheric CO
2
are not tracking changes in human
emissions.
(7) On the time scale investigated, the overriding effect of large vol-
canic eruptions appears to be a reduction of atmospheric CO
2
,
presumably due to the dominance of associated cooling effects
from clouds associated with volcanic gases/aerosols and volca-
nic debris.
(8) Since at least 1980 changes in global temperature, and presum-
ably especially southern ocean temperature, appear to represent
a major control on changes in atmospheric CO
2
.
Acknowledgments
The present study could not have been carried out without the ex-
istence of open-access international databases on temperature and
atmospheric CO
2
. We thank one unknown referee for pointing out
the variable correlation between changes in the release of CO
2
from
anthropogene sources and changes in atmospheric CO
2
on a decadal
time scale. In addition, this referee suggested the hemispheric analy-
ses (Section 7). We are nally grateful for informal discussions in Oslo
with Drs. O.H. Ellestad, O. Engvold, and P. Brekke. Also Drs. T.V.
Segalstad and H. Yndestad have for long time been important sources
of information and inspiration.
References
Alley, R.B., Clarck, P.U., 1999. The deglaciation of the northern hemisphere: a global
perspective. Annual Earth Planetary Science 27, 149182.
Bacastow, R.B., 1976. Modulation of atmospheric carbon dioxide by the Southern Oscillation.
Nature 261, 116118.
Caillon, N., Severinghaus, J.P., Jouzel, J., Barnola, J.-M., Kang, J., Lipenkov, V.Y., 2003.
Timing of atmospheric CO
2
and Antarctic temperature changes across Termination
III. Science 299 (5613), 17281731.
Conway, T.J., Tans, P.P., Waterman, L.S., Thoning, K.W., Kitzis, D.R., Masarie, K.A.,
Zhang, N., 1994. Evidence for interannual variability of the carbon cycle from
the NOAA/CMDL global air sampling network. Journal of Geophysical Research
99, 2283122855.
Essenhigh, R.E., 2009. Potential dependence of global warming on the residence time
(RT) in the atmosphere of anthropogenically sourced carbon dioxide. Energy &
Fuels 23, 27732784.
Goldberg, F., 2008. Rate of increasing concentrations of atmospheric Carbon Dioxide con-
trolled by natural temperature variations. Energy & Environment 19 (7), 6777.
Humlum, O., Solheim, J.-E., Stordahl, K., 2011. Identifying natural contributions to late
Holocene climate change. Global and Planetary Change 79, 145156. http://
dx.doi.org/10.1016/j.gloplacha.2011.09.005.
Humlum, O., Solheim, J.-E., Stordahl, K., 2012. Spectral analysis of the Svalbard temper-
ature record 19122010. Advances in Meteorology 2012. http://dx.doi.org/
10.1155/2012/175296 (Article ID 175296, 14 pages).
IPCC AR4, 2007. Contribution of Working Group I to the Fourth Assessment Report of
the Intergovernmental Panel on Climate Change. Cambridge University Press,
Cambridge, United Kingdom and New York, NY, USA.
Jaworowski, Z., Segalstad, T.V., Hisdal, V., 1992. Atmospheric CO
2
and Global Warming:
A Critical Review, 2nd revised edition. : Letters, Vol. 119. Norwegian Polar Institute
(76 pp.).
Keeling, C.D., Whorf, T.P., Wahlen, M., van der Plicht, J., 1995. Interannual extremes in the
rate of rise of atmospheric carbon dioxide since 1980. Nature 375, 666670.
Lamb, H.H., 1972. Climate: Present, Past and Future. : Fundamentals and Climate Now,
Vol. 1. Methuen & Co Ltd., London (613 pp. SBN 416-11530-6).
Langenfelds, R.L., Francey, R.J., Pak, B.C., Steele, L.P., Lloyd, J., Trudinger, C.M., Allison, C.E.,
2002. Interannual growth rate variations of atmospheric CO
2
and its delta C-13, H-2,
CH4, and CO between 1992 and 1999 linked to biomass burning. Global Biogeochem-
ical Cycles 16. http://dx.doi.org/10.1029/2001GB001466.
Lockwood, M., Fröhlich, C., 2008. Recent oppositely directed trends in solar climate
forcings and the global mean surface air temperature. II. Different reconstruc-
tions of the total solar irradiance variation and dependence on response time
scale. Proceedings of the Royal Society A 464, 13671385. http://dx.doi.org/
10.1098/ispa.2007.0347.
Lorius, C., Raynaud, D., Jouzel, J., Hansen, J., Le Treut, H., 1990. The ice-core recordclimate
sensitivity and future greenhouse warming. Natur e 347 (6289), 139145.
Lüthi, D., Le Floch, M., Bereiter, B., Blunier, T., Barnola, J.-M., Siegenthaler, U., Raynaud,
D., Jouzel, J., Fischer, H., Kawamura, K., Stocker, T.F., 2008. High-resolution carbon
dioxide concentration record 650,000800,000 years before present. Nature 453,
379382. http://dx.doi.org/10.1038/nature06949.
Martin, P., Archer, D., Lea, D.W., 2005. Role of deep sea temperature in the carbon cycle
during the last glacial. Paleoceanography 20 (2), 110.
Masarie, K.A., Tans, P.P., 1995. Extension and integration of atmospheric carbon dioxide
data into a globally consistent measurement record. Journal of Geophysical Re-
search 100, 1159311610.
Monnin, E., Indermühle, A., Dällenbach, A., Flückiger, J., Stauffer, B., Stocker, T.F.,
Raynaud, D., Barnola, J.-M., 2001. Atmospheric CO
2
concentrations over the last
glacial termination. Science 291, 112114.
Mudelsee, M., 2001. The phase relations among atmospheric CO
2
content, temperature
and globalice volume over the past420 ka. Quaternary Science Reviews 20, 583589.
Nakazawa, T., Moromoto, S., Aoki, S., Tanaka, M., 1997. Temporal and spatial variations
of the carbon isotopic ratio of atmospheric carbon dioxide in the Western Pacic
region. Journal of Geophysical Research 102, 12711285.
Newhall, C.G., Self, S., 1982. The volcanic explosivity index (VEI): an estimate of explo-
sive magnitude for historical volcanism. Journal of Geophysical Research-Oceans &
Atmospheres 87, 12311238.
Njau, E.C., 2007. Formulations of human-induced variations in global temperature. In-
ternational Journal of Renewable Energy 32, 22112222.
Scafetta, N., 2011. Testing an astronomically based decadal-scale empirical harmonic
climate model versus the IPCC (2007) general circulation climate models. Journal
of Atmospheric and Solar-Terrestrial Physics. http://dx.doi.org/10.1016/j.jastp.
2011.12.005.
SeaSolve, 2003. AutoSignal. Pioneering Automated Signal Analysis and Advanced Fil-
tering. User's Manual. SeaSolve Software Inc., Framingham, MA, USA.
Segalstad, T.V., 1998. Carbon Cycle Modelling and the Residence Time of Natural and
Anthropogenic Atmospheric CO
2
: On the Construction of the Greenhouse Effect
68 O. Humlum et al. / Global and Planetary Change 100 (2013) 5169
Author's personal copy
Global WarmingDogma. In: Bate, R. (Ed.), Global Warming: The Continuing De-
bate. ESEF, Cambridge, U.K. ISBN: 0952773422, pp. 184219.
Severinghaus, J.P., Sowers, T., Brook, E.J., Alley, R.B., Bender, M.L., 1998. Timing of abrupt
climate change at the end of the Younger Dryas interval from thermally fractionated
gases in polar ice. Nature 391, 141146. http://dx.doi.org/10.1038/34346.
Shackleton, N.J., 2000. The 100,000 year ice-age cycle identied and found to lag tem-
perature, carbon dioxide and orbital eccentricity. Science 289, 18971902.
Shakun, J.D., Clark, P.U., He, F., Marcott, S.A., Mix, A.C., Liu, Z., Otto-Bliesner, B.,
Schmittner, A., Bard, E., 2012. Global warming preceded by increasing carbon diox-
ide concentrations during the last deglaciation. Nature 484, 4955. http://
dx.doi.org/10.1038/nature10915.
Simkin, T., Siebert, L., 1994. Volcanoes of the World, 2nd edition. Geoscience Press in
association with the Smithsonian Institution Global Volcanism Program, Tucson
AZ . (368 pp.).
Solanki, S.K., Usoskin, I.G., Kromer, B., Schüssler, M., Beer, J., 2004. Unusual activity of
the Sun during recent decades compared to the previous 11,000 years. Nature
431, 10841087.
Stauning, P., 2011. Solar activityclimate relations: a different approach. Journal of At-
mospheric and Solar-Terrestrial Physics 73, 19992012.
Thompson, D.W.T., Wallace, J.M., Jones, P.D., Kennedy, J.J., 2009. Identifying signatures
of natural climate variability in time series of global-mean surface temperature.
Methodology and insights. Journal of Climate 22, 61206141. http://dx.doi.org/
10.1175/2009JCLI3089.1.
Toggweiler, J.R., 1999. Variation of atmospheric CO
2
by ventilation of the ocean's
deepest water. Paleoceanography 14 (5), 571588.
Toggweiler, J.R., Lea, D.W., 2010. Temperature differences between the hemispheres
and ice age climate variability. Palaeoceanography 25, PA2212.
Web references
CO
2
globally averaged marine surface data (accessed March 11, 2012): ftp://ftp.cmdl.
noaa.gov/ccg/co2/trends/co2_mm_gl.txt
HadCRUT3 monthly global surface air temperature (accessed March 11, 2012): http://
www.metofce.gov.uk/hadobs/hadcrut3/diagnostics/global/nh+sh/monthly
HadSST2 monthly global sea surface temperature (accessed March 11, 2012): http://
www.cru.uea.ac.uk/cru/data/temperature/hadsst2gl.txt
UAH monthly global lower troposphere temperature (accessed March 11, 2012):
http://vortex.nsstc.uah.edu/data/msu/t2lt/uahncdc.lt
Goddard Institute for Space Studies, GISS (accessed March 11, 2012): http://data.giss.
nasa.gov/gistemp/tabledata_v3/GLB.Ts+dSST.txt
National Climatic Data Center, NCDC (accessed March 11, 2012): http://www.ncdc.
noaa.gov/cmb-faq/anomalies.phpIndividual station atmospheric CO
2
records from
Earth System Research Laboratory, NOAA (accessed July 25, 2012): http://www.esrl.
noaa.gov/gmd/ccgg/iadv/
Carbon Dioxide Information Analysis Center, CDIAC (accessed March 11, 2012): http://
cdiac.ornl.gov/epubs/fossil_fuel_CO2_emissions_gridded_monthly_v2011.html
Global Volcanism Program, GWP (accessed March 11, 2012): http://www.volcano.
si.edu/world/largeeruptions.cfm
69O. Humlum et al. / Global and Planetary Change 100 (2013) 5169
... Strong coupling of anomalous CO2 to tropical temperature was previously noted by Humlum et al. (2013). They observed further that anomalous CO2 appeared originally in the tropics but then advanced poleward into each hemisphere. ...
... The lag required by the conservation law is thus  8.5 months. It is close to the observed lag that emerges from the cross covariance between CO2 and temperature (Humlum et al. 2013;Salby 2013). (2) ...
... Because it applies irrespective of time scale (), this consequence of the conservation law holds equally for interannual fluctuations and the longterm increase of net emission. As observed net emission varies in phase with anomalous temperature, on all relevant time scales (Fig. 5b), anomalous CO2 must also lag anomalous temperature by a quarter cycle -as is, in fact, observed (Humlum et al. 2013;Salby 2013). ...
Article
Full-text available
Unlike much of the Earth, surface temperature in the tropics underwent a systematic and sustained increase during the satellite era. Due to the temperature dependence of surface processes which regulate CO2 emission, that long-term change should exert a similar influence on atmospheric CO2. We develop how this influence would manifest in the evolution of CO2. Observed records are then used to investigate the interdependence of surface temperature and net CO2 emission-the component of emission that actually changes CO2. Thermally-induced emission, especially from tropical land surface, is found to represent much of the observed evolution of net CO2 emission. It accounts for sporadic intensifications of net emission that operate on interannual time scales, notably, during episodes of El Niño. Accounted for equally well is the long-term intensification of net emission during the last half century. Jointly, these unsteady components of net emission determine the thermally-induced component of anomalous CO2. It closely tracks the observed evolution of CO2.
... The change in temperature changed the rate of CO 2 growth after 5 months. Humlum et al. 11 compared changes in his CO 2 concentration and land-sea surface temperature over the period from January 1980 to December 2011, during which time they examined the phase relationship between CO 2 and temperature, and found that changes in CO 2 always lag changes in temperature by 10 to 12 months. Our case is consistent with the results of Humlum et al. ...
... Since the decomposition of plants is a biological process, it does not change immediately when the temperature changes, and there is some time lag. In this paper, a time lag of 5 months was obtained between changes in global temperature and the rate of CO 2 increase, while a time lag of 10-12 months was also reported for changes in temperature and CO 2 concentration by Humlum et al. 11 . ...
Preprint
Full-text available
We compared satellite-based UAH Earth surface temperature, NOAA ENSO index, and Earth-surface CO 2 change rate in detail. These three variables were well correlated over the observed 40 years. Temperature changed about one year after ENSO index changed, and CO 2 change rate followed temperature change by several months. CO 2 emission and absorption at the Earth’s surface responded to temperature changes. It was like Earth’s breathing. CO 2 emission due to higher temperature was related to plant respiration or decomposition processes.
... From faster oscillations on the 14 CO2 decay (SH1) and a cross-correlation analysis of interannual CO2 and temperature fluctuations (Humlum et al. 2013;Salby 2013) it even follows that the direct absorption time  can be as short as about 1 yr, which under respective conditions also ties down the effective absorption time. ...
... For this uptake and release of CO2 it is reasonable to distinguish between direct absorption processes on a timescale even as short as about one year, as this follows from observations of the faster oscillations on the 14 CO2 decay (SH1) and also from cross-correlation analyses of interannual CO2 and temperature fluctuations (Humlum et al. 2013;Salby 2013), and on the other hand between an effective absorption on an extended timescale, which is determined by re-emission from extraneous reservoirs. This re-emission is assumed to be proportional to a previous absorption by these reservoirs and represents a good approximation for a coupled balance scheme of adjacent reservoirs with the atmosphere. ...
Article
Full-text available
The carbon cycle is of fundamental importance to estimate the influence of anthropogenic emissions on the atmospheric CO2 concentration, and thus, to classify the impact of these emissions on global warming. Different models have been developed, which under simplified assumptions can well reproduce the observed CO2 concentration over recent years, but they also lead to contradictory interpretations of the human impact. Here we consider, how far such suppositions are substantiated or must be made responsible for significant misinterpretations. We present detailed own calculations based on the Conservation Law, which reproduce all details of the measured atmospheric CO2 concentration over the Mauna Loa Era. In particular, they allow to deduce an upper limit of 35% for the anthropogenic contribution to the observed increase of CO2 over the Mauna Loa Era, and a more likely value of 14%. Under non-equilibrium conditions between the Earth's surface and troposphere this even gives a lower bound of only 3.5%. The importance of only one unitary time scale for the removal of anthropogenic and natural CO2 emissions from the atmosphere, characterized by an effective absorption time, is discussed.
... Through the conservation law, the strong correspondence to observed net emission follows theoretically from behavior in the tropics -not in the extratropics, where anthropogenic emission is concentrated. Independently, the same correspondence to observed Enet follows empirically from anomalous temperature in the tropics (SH2), as well as from a time-lag analysis of anomalous CO2 (Humlum et al. 2013). Especially noteworthy is that, in the time-lag analysis, strong coupling of anomalous CO2 appears initially in the tropics, but then advances poleward into each hemisphere. ...
... According to (50), it experiences effective an effective absorption time of less than a year.It is noteworthy that absorption of the HP component of anomalous CO2 can also be deduced empirically. Interannual fluctuations of CO2 are observed to operate coherently with but lag interannual fluctuations of temperature(Humlum et al. 2013;Salby 2013;Harde 2019). Independent of the above analysis, the observed lag reveals effective absorption balance (43) is now solved numerically for thermally-induced net emission, Enet(T), through 4th order Runge-Kutta treatment. ...
Article
Full-text available
Unlike elsewhere on the globe, temperature in the tropics has increased systematically. From observed tropical temperature, numerical simulations have reproduced the observed evolution of atmospheric CO2, including its annual cycle. Much the same has followed empirically from the observed covariance between tropical temperature and net emission of CO2 - the component of emission that actually changes CO2. Intensified over tropical land, the robust coherence between those observed features establishes that the changes of tropical temperature do not follow from changes of CO2, but rather produce them. Here, we investigate the physical mechanisms through which observed warming in the tropics can produce the observed evolution of CO2. The conservation law governing atmospheric CO2, supported by the observed temperature dependence of surface fluxes and observed temperature in the tropics, is used to calculate the time-varying response of CO2 net emission. Thermally-induced emission is found to account for the preponderance of CO2 net emission and, thereby, for the vast majority of anomalous CO2. Magnified over tropical land, it represents interannual intensifications of net emission, notably during episodes of El Niño. Represented equally well is the long-term intensification of net emission during the last half century. That intensification alone accounts for about half of the observed increase of atmospheric CO2. Anomalous CO2 introduced by the calculated net emission closely tracks the observed evolution of atmospheric CO2. Having similar dependence on temperature is methane. Like net emission of CO2, net emission of methane is intensified by tropical warming. The simultaneous intensification of CO2 and CH4 emission provides a unified understanding of their joint increase, one that follows naturally from thermally-induced emission.
... MD analyzed only GCM simulations that implemented the RCP8.5 high-emission scenario. As for individual GCMs, increases in CO 2 concentrations and temperature are roughly correlated (Humlum et al., 2013). We also applied the PT-MA approach for other RCPs. ...
Article
Full-text available
Almost no hydrological model takes into account that changes in evapotranspiration are affected by how vegetation responds to changing CO2 and climate. This severely limits their ability to quantify the impact of climate change on evapotranspiration and, thus, water resources. As the simulation of vegetation responses is both complex and very uncertain, we recommend a simple approach to considering (in climate change impact studies with hydrological models) the uncertainty that the vegetation response causes with respect to the estimation of future potential evapotranspiration (PET). To quantify this uncertainty in a simple manner, we propose running the hydrological model in two variants: with its standard PET approach and with a modified approach to compute PET. In the case of PET equations containing stomatal conductance, the modified approach can be implemented by adjusting the conductance. We introduce a modified approach for hydrological models that computes PET as a function of net radiation and temperature only, i.e., with the Priestley–Taylor (PT) equation. The new PT-MA approach is based on the work of Milly and Dunne (2016) (MD), who compared the change in non-water-stressed actual evapotranspiration (NWSAET) as computed by an ensemble of global climate models (GCMs), which simulate vegetation response as well as interactions between the atmosphere and the land surface, with various methods to compute PET change. Based on this comparison, MD proposed estimating the impact of climate change on PET as a function of only the change in net energy input at the land surface. PT-MA retains the impact of temperature on daily to interannual as well as spatial PET variations but removes the impact of the long-term temperature trend on PET such that long-term changes in future PET are driven by changes in net radiation only. We implemented PT-MA in the global hydrological model WaterGAP 2.2d and computed daily time series of PET between 1901 and 2099 using the bias-adjusted output of four GCMs. Increases in GCM-derived NWSAET between the end of the 20th and the end of the 21st century for Representative Concentration Pathway 8.5 (RCP8.5) are simulated well by WaterGAP if PT-MA is applied but are severely overestimated with the standard PT method. Application of PT-MA in WaterGAP results in smaller future decreases or larger future increases in renewable water resources (expressed as the variable RWR) compared with the standard PT method, except in a small number of grid cells where increased inflow from upstream areas due to increased upstream runoff leads to enhanced evapotranspiration from surface water bodies or irrigated fields. On about 20 % of the global land area, PT-MA leads to an increase in RWR that is more than 20 % higher than in the case of standard PT, while on more than 10 % of the global land area, the projected RWR decrease is reduced by more than 20 %. While the modified approach to compute PET is likely to avoid the overestimation of future drying in many if not most regions, the vegetation response in other regions may be such that the application of the standard PET leads to more likely changes in PET. As these regions cannot be identified with certainty, the proposed ensemble approach with two hydrological model variants serves to represent the uncertainty in hydrological changes due to the vegetation response to climate change that is not represented in the model.
... The atmospheric carbon dioxide from about 20% to very low during the succeeding Cambrian period. Carbon dioxide is unusual in becoming more soluble in water at lower temperatures [13]. Conversely, it is degassed when the seas warm up. ...
Article
Full-text available
Summarizes the potential consequences of removal of atmospheric carbon dioxide from the air.
... Since the decomposition of plants is a biological process, it does not change immediately when the temperature changes, and there is some time lag. In this paper, a time lag of 5 months was obtained between changes in global temperature and the rate of CO 2 increase, while a time lag of 10-12 months was also reported for changes in temperature and CO 2 concentration by Humlum et al. 11 . ...
Preprint
Full-text available
We compared satellite-based UAH Earth surface temperature, NOAA ENSO index, and Earth-surface CO 2 change rate in detail. These three variables were well correlated over the observed 40 years. The temperature changed approximately one year after the ENSO index changed, and the CO 2 change rate followed the temperature change by several months. CO 2 emission and absorption at the Earth’s surface responded to temperature changes. It was like Earth’s breathing. CO 2 emission due to higher temperature was related to plant respiration or decomposition processes.
... Altogether we find a strong correspondence to observed net emission that follows theoretically from behavior in the tropics -not in the extratropics, where anthropogenic emission is concentrated. Independently, the same correspondence to observed net emission follows empirically from anomalous temperature in the tropics (see Part II), as well as from a time-lag analysis of anomalous CO2 [11]. ...
Article
Full-text available
The Intergovernmental Panel on Climate Change (IPCC) classifies the human influence on our climate as extremely likely to be the main reason of global warming over the last decades. Particularly anthropogenic emissions of carbon compounds, with carbon dioxide (CO2) as the main culprit and methane (CH4) as a distant second, are made responsible for the observed temperature changes, while any natural forcings are almost completely excluded. This post summarizes the results of three studies [1-3] addressing the question, how much human or native emissions can be made responsible for the observed increase of Greenhouse Gases (GHG), in particular the rising mixing ratio of CO2 in the atmosphere.
... In Section 2 we have already discussed that such an approach violates the Equivalence Principle and is also in contradiction to observations of the 14 C-decay, from which we derive an absorption time shorter than 10 yrs. Direct absorption processes can even be as short as one year, as this follows from the faster oscillations on the 14 CO2 decay (Salby & Harde 2021) and also from cross-correlation analyses of interannual CO2 and temperature fluctuations (Humlum et al. 2013;Salby 2013). ...
Article
Full-text available
This reply refutes all misstatements, that were published by F. Engelbeen as Comment on an article "Understanding Increasing Atmospheric CO2 " by Hermann Harde.
... (i) that the well documented delayed and pure native emissions of CO2 and methane (CH4) to sea and air temperature changes (see, e.g., Petit et al. [3]; Monnin et al. [4]; Caillon et al. [5]; Torn & Harte [6]; Humlum et al. [7]; Salby [8]) are not further considered in AR5 or AR6 with their consequences for interpreting actual climate changes (for a detailed con-sideration of the "hen-or-egg causality" see also Koutsoyiannis & Kundzewicz [9]); ...
Article
Full-text available
The Intergovernmental Panel on Climate Change classifies the human influence on our climate as extremely likely to be the main reason of global warming over the last decades. Particularly anthropogenic emissions of carbon dioxide are made responsible for the observed temperature changes, while any natural forcings are almost completely excluded. However, detailed own calculations with an advanced energy-radiation-balance model indicate that the temperature increase and its variations over the last 140 years can much better be explained by additionally including solar radiative forcing and its amplification by induced cloud cover changes. We present simulations based on different time series of the total solar irradiance and compare them with composed land-ocean-surface temperature measurements of the Northern Hemisphere. From these simulations we follow that CO2 should not have contributed more than about one third to global warming over the last century, while solar variations over this period can well explain two thirds of the increase.
Article
Full-text available
The most important basis of the hypothesis of man-made climatic warming due to burning of fossil carbon fuels are the measurements of CO2 in air and in glacier ice, hydrogen and oxygen isotopes in glacier ice, carbon isotopes in tree rings, and 100-150 yr long atmospheric temperature records. In this paper we critically review these measurements and their interpretations, in order to test the nowadays widely accepted postulate that "the change in atmospheric CO2 is not just a fluctuation of nature, but is predominantly the consequence of the activities of mankind - chiefly the burning of fossil fuels such as coal, gas and oil' (Rowland and Isaksen, 1988). We also compare the quantities of anthropogenic contributions with known fluxes of natural reservoirs, and discuss air temperature and glacier balance records, which should reveal signals of an increasing "greenhouse effect'. The above discussion supports the opinion (Wigley et al., 1989) that the anthropogenic increasing "greenhouse' warming signal has not yet been detected in a rigorous way. -from Authors
Article
Full-text available
Rapid temperature change fractionates gas isotopes in unconsolidated snow, producing a signal that is preserved in trapped air bubbles as the snow forms ice. The fractionation of nitrogen and argon isotopes at the end of the Younger Dryas cold interval, recorded in Greenland ice, demonstrates that warming at this time was abrupt. This warming coincides with the onset of a prominent rise in atmospheric methane concentration, indicating that the climate change was synchronous (within a few decades) over a region of at least hemispheric extent, and providing constraints on previously proposed mechanisms of climate change at this time. The depth of the nitrogen-isotope signal relative to the depth of the climate change recorded in the ice matrix indicates that, during the Younger Dryas, the summit of Greenland was 15 +/- 3°C colder than today.
Article
Full-text available
A composite estimate of the magnitude of past explosive eruptions, termed the volcanic Explosivity Index (VEI), is proposed as a semiquantitative compromise between poor data and the need in various disciplines to evaluate the record of past volcanism. The VEI has been assigned to over 8000 historic and prehistoric eruptions, and a complete list is available in a companion document.-from Authors
Article
Full-text available
Climate development with possible anthropogenic effects occurs on a background of natural climatic variations, which may be considerable, and especially in the Arctic. Natural climate variations however remain poorly understood, although they remain important for discriminating between natural and anthropogenic influences on current climate change. Using the Svalbard (78°N) surface air temperature record 1912–2010 as an example, we here suggest a road ahead to identify and describe such natural climate variations. By means of Fourier and wavelet analysis the record is decomposed into time-frequency space, to extract information on periodic signals and their amplitude and variation over time. By this we identify several cyclic variations on the time scale investigated. These oscillations are present in the annual record, as well as in seasonal subsets of the record. Using only three oscillations it is possible to hindcast the Svalbard temperature record well. We suggest that such persistent oscillations may be used for forecasting the overall features of future temperature changes for a limited period, about 10–25% of the record length. Our main focus is on identifying the character of recurrent natural temperature variations, but we also comment briefly on possible physical explanations for some of the identified cyclic variations.
Article
Full-text available
Global-mean surface temperature is affected by both natural variability and anthropogenic forcing. This study is concerned with identifying and removing from global-mean temperatures the signatures of natural climate variability over the period January 1900-March 2009. A series of simple, physically based methodologies are developed and applied to isolate the climate impacts of three known sources of natural variability: the El Nino-Southern Oscillation (ENSO), variations in the advection of marine air masses over the high-latitude continents during winter, and aerosols injected into the stratosphere by explosive volcanic eruptions. After the effects of ENSO and high-latitude temperature advection are removed from the global-mean temperature record, the signatures of volcanic eruptions and changes in instrumentation become more clearly apparent. After the volcanic eruptions are subsequently filtered from the record, the residual time series reveals a nearly monotonic global warming pattern since similar to 1950. The results also reveal coupling between the land and ocean areas on the interannual time scale that transcends the effects of ENSO and volcanic eruptions. Globally averaged land and ocean temperatures are most strongly correlated when ocean leads land by; 2-3 months. These coupled fluctuations exhibit a complicated spatial signature with largest-amplitude sea surface temperature perturbations over the Atlantic Ocean.
Article
Studies of the temperature anomalies during the last 27 years show a close relationship with the varying increase of CO 2 in the atmosphere. Volcanic eruptions and La Niñas reduce CO 2 values and El Niños increase them. This close relationship strongly indicates that ocean temperatures and the solubility of CO 2 in seawater control the amount of CO 2 being absorbed or released by the oceans. It is therefore likely that the increased CO 2 concentration in the atmosphere is due to a natural global warming and that CO 2 produced through fossil fuel combustion by humans can not disrupt this balance. An advanced statistical multiregression analysis confirms this conclusion. Therefore it is likely that there is no anthropogenic climate change on a global scale. The natural exchange of CO 2 between ocean, biomass on land and the atmosphere is very large. In only four to five years all the CO 2 in the atmosphere has been recycled through the oceans and the biomass system. The annual anthropogenic human production of CO 2 is neutralized by nature in as little as 12 days. Recent studies of the solar forcing, changes in cosmic radiation and its role in cloud formations explain the global warming that has taken place since 1910.