ArticlePDF Available

European Seasonal and Annual Temperature Variability, Trends, and Extremes Since 1500

Authors:

Abstract and Figures

Multiproxy reconstructions of monthly and seasonal surface temperature fields for Europe back to 1500 show that the late 20th- and early 21st-century European climate is very likely (>95% confidence level) warmer than that of any time during the past 500 years. This agrees with findings for the entire Northern Hemisphere. European winter average temperatures during the period 1500 to 1900 were reduced by approximately 0.5 degrees C (0.25 degrees C for annual mean temperatures) compared to the 20th century. Summer temperatures did not experience systematic century-scale cooling relative to present conditions. The coldest European winter was 1708/1709; 2003 was by far the hottest summer.
Content may be subject to copyright.
0.12 for this grain is higher than any reported
to date in presolar oxide grains (30, 33, 34).
However, shell H-burning and third dredge-up
in low-mass AGB stars produce a
26
Al/
27
Al
ratio of only 1.5 10
3
in the envelope (26).
Another consideration is that the grain has to
form before dredge-up of C turns the star into a
carbon star (23). Consequently, another process
must be invoked to account for the large
26
Al/
27
Al. Previously, CBP was invoked to account
for the high
26
Al/
27
Al ratios found in a corun-
dum and a hibonite grain (34). As mentioned
already, CBP occurs in low-mass stars during
the RGB and the thermally pulsing AGB phas-
es. Although CBP alters the O and Al isotopic
abundances simultaneously, there is no correla-
tion between the two systems because the de-
crease in
18
O/
16
O is dependent on the rate of
mass circulation (M
˙
), whereas
26
Al/
27
Al de-
pends on the maximum temperature of the cir-
culated material (T
P
), which in turn depends on
depth of penetration (23). According to CBP
models for a 1.5 M
J
(solar mass) star (23),
26
Al
is abundantly produced for log T
P
7.65, but
the large
26
Al/
27
Al observed in the presolar
silicate grain requires log T
P
7.76 and very
deep mixing. In addition, the O isotopic com-
position of this grain requires M
˙
10
6
M
J
per year. These values for T
P
and M
˙
are feasible
in low-mass thermally pulsing AGB stars un-
dergoing CBP. It is apparent that the combined
isotopic analysis of O and Mg provides more
detailed information on deep mixing processes
than just one isotopic system would provide.
In combining the data for all presolar silicates
in IDPs (11, 12), we arrive at a total abundance of
890 ppm (9). In contrast, the abundance in
Acfer 094 relative to the matrix material is 40
ppm. This difference confirms that IDPs are more
primitive than any meteorite. The abundance of
circumstellar silicates in the matrix of the ordinary
chondrites Semarkona and Bishunpur has been
inferred to be 15 ppm (14, 15) compared with
40 ppm in Acfer 094. These ordinary chondrites
have undergone more aqueous alteration than
Acfer 094 and formed at higher temperatures.
This difference in abundance gives us a first look
at the effects of meteorite formation histories and
parent body processes on presolar grain survival.
Determination of the abundances of presolar sili-
cates in other meteorite classes will give us a
means of studying the physical conditions in dif-
ferent solar system environments. Due to the un-
usually primitive nature of Acfer 094, we do not
expect the abundance to be higher in other mete-
orites, but we still do not understand the destruc-
tive processes affecting presolar silicate grains
well enough to make this a firm prediction.
References and Notes
1. E. Zinner, in Meteorites, Comets, and Planets, A. M.
Davis, Ed. (Elsevier, Oxford, UK, 2004), vol. 1, pp. 17–39.
2. C. Waelkens et al., Astron. Astrophys. 315, L245 (1996).
3. K. Malfait et al., Astron. Astrophys. 332, L25 (1998).
4. L. B. F. M. Waters et al., Astron. Astrophys. 315, L361
(1996).
5. K. Demyk, E. Dartois, H. Wiesemeyer, A. P. Jones, L.
d’Hendecourt, Astron. Astrophys. 364, 170 (2000).
6. L. R. Nittler, thesis, Washington University (1996).
7. S. Messenger, T. J. Bernatowicz, Meteorit. Planet. Sci.
35, A109 (2000).
8. C. M. O’D. Alexander, L. Nittler, F. Tera, Lunar Planet.
Sci. XXXII, A2191 (2001).
9. Materials and methods are available as supporting
material on Science Online.
10. A. Nguyen, E. Zinner, R. S. Lewis, Publ. Astron. Soc.
Aust. 20, 382 (2003).
11. S. Messenger, L. P. Keller, F. J. Stadermann, R. M.
Walker, E. Zinner, Science 300, 105 (2003).
12. C. Floss, F. J. Stadermann, Lunar Planet. Sci. XXXV,
A1281 (2004).
13. J. P. Bradley, Science 265, 925 (1994).
14. S. Mostefaoui, P. Hoppe, K. K. Marhas, E. Gro¨ner,
Meteorit. Planet. Sci. 38, A99 (2003).
15. S. Mostefaoui, K. K. Marhas, P. Hoppe, Lunar Planet.
Sci. XXXV, A1593 (2004).
16. F. J. Molster, L. B. F. M. Waters, A. G. G. M. Tielens, C.
Koike, H. Chihara, Astron. Astrophys. 382, 241 (2002).
17. F. J. Molster et al., Nature 401, 563 (1999).
18. S. Weinbruch, H. Palme, W. F. Mu¨ller, A. El Goresy,
Meteoritics 25, 115 (1990).
19. X. Hua, J. Adam, A. El Goresy, H. Palme, Geochim.
Cosmochim. Acta 52, 1389 (1988).
20. E. Zinner, Annu. Rev. Earth Planet. Sci. 26, 147 (1998).
21. A. I. Boothroyd, I.-J. Sackmann, Astrophys. J. 510,232
(1999).
22. G. J. Wasserburg, A. I. Boothroyd, I.-J. Sackmann,
Astrophys. J. 447, L37 (1995).
23. K. M. Nollett, M. Busso, G. J. Wasserburg, Astrophys.
J. 582, 1036 (2003).
24. A. I. Boothroyd, I.-J. Sackmann, G. J. Wasserburg,
Astrophys. J. 442, L21 (1995).
25. R. C. Cannon, C. A. Frost, J. C. Lattanzio, P. R. Wood,
in Nuclei in the Cosmos III, M. Busso, R. Gallino, C. M.
Raiteri, Eds. (AIP, New York, 1995), pp. 469 472.
26. M. Forestini, G. Paulus, M. Arnould, Astron. Astrophys.
252, 597 (1991).
27. N. Mowlavi, G. Meynet, Astron. Astrophys. 361, 959
(2000).
28. A. I. Karakas, J. C. Lattanzio, Publ. Astron. Soc. Aust.
20, 279 (2003).
29. M. Busso, R. Gallino, G. J. Wasserburg, Annu. Rev.
Astron. Astrophys. 37, 239 (1999).
30. L. R. Nittler, C. M. O’D. Alexander, X. Gao, R. M.
Walker, E. Zinner, Astrophys. J. 483, 475 (1997).
31. C. M. O’D. Alexander, L. R. Nittler, Astrophys. J. 519,
222 (1999).
32. B.-G. Choi, G. R. Huss, G. J. Wasserburg, R. Gallino,
Science 282, 1284 (1998).
33. I. D. Hutcheon, G. R. Huss, A. J. Fahey, G. J. Wasser-
burg, Astrophys. J. 425, L97 (1994).
34. B.-G. Choi, G. J. Wasserburg, G. R. Huss, Astrophys. J.
522, L133 (1999).
35. We thank the reviewer for valuable comments and are
grateful to S. Amari and C. Floss for their assistance with
SEM measurements and helpful discussions. We also thank
S. Messenger and F. Stadermann for developing the soft-
ware for processing the isotopic images. Supported by
NASA grants NAG5-10426 and NAG5-11545.
Supporting Online Material
www.sciencemag.org/cgi/content/full/303/5663/1496/
DC1
Materials and Methods
Table S1
References
5 December 2003; accepted 30 January 2004
European Seasonal and Annual
Temperature Variability, Trends,
and Extremes Since 1500
Ju¨rg Luterbacher,
1,2
* Daniel Dietrich,
3
Elena Xoplaki,
2
Martin
Grosjean,
1
Heinz Wanner
1,2
Multiproxy reconstructions of monthly and seasonal surface temperature fields
for Europe back to 1500 show that the late 20th- and early 21st-century
European climate is very likely (95% confidence level) warmer than that of
any time during the past 500 years. This agrees with findings for the entire
Northern Hemisphere. European winter average temperatures during the period
1500 to 1900 were reduced by 0.5°C (0.25°C for annual mean temperatures)
compared to the 20th century. Summer temperatures did not experience sys-
tematic century-scale cooling relative to present conditions. The coldest Eu-
ropean winter was 1708/1709; 2003 was by far the hottest summer.
Detailed insight into high-resolution temporal
and spatial patterns of climate change during
previous centuries is essential for assessing the
degree to which late 20th-century changes may
be unusual in the light of preindustrial natural
climate variability (1–3). Climate change at sea-
sonal to annual resolutions for recent centuries
has been highlighted in a number of studies,
which have included climate modeling experi-
ments with estimated natural and anthropogenic
radiative-forcing changes (46) and empirical
hemispheric or global reconstructions. Such re-
constructions are based either on natural ar-
chives only (such as ice cores, tree rings, spe-
leothems, varved sediments, and subsurface
temperature profiles obtained from borehole
measurements) or on multiproxy networks that
amalgamate natural proxy indicators with cli-
mate information obtained from early instru-
mental and documentary evidence (714). A
number of these reconstructions support the
conclusion that the warmth of the late 20th
century is likely unprecedented in the Northern
1
National Center of Competence in Research (NCCR) Cli-
mate,
2
Institute of Geography, Climatology, and Meteorol-
ogy,
3
Department of Mathematical Statistics and Actuarial
Science, University of Bern, CH–3012 Bern, Switzerland.
*To whom correspondence should be addressed. E-
mail: juerg@giub.unibe.ch
R EPORTS
www.sciencemag.org SCIENCE VOL 303 5 MARCH 2004 1499
Hemisphere in the past 1000 years and cannot
be explained by natural forcings alone (15).
Hemispheric and global temperature recon-
structions do not provide information about re-
gional-scale variations, such as the intrinsic sea-
sonal patterns of climate change as they have
occurred in Europe during the past centuries.
The few European-scale temperature reconstruc-
tions (7, 1619) have revealed information for the
winter or summer half-year or for annual to mul-
tiannual mean values. Changes in the full annual
cycle have typically not been addressed, because
of the limited year-round information provided by
most natural climate proxy data (2, 20).
Regional and temporal high-resolution re-
constructions also illuminate key climatic fea-
tures, such as regionally very hot or cool sum-
mers and very mild or cold winters, that may be
masked in a hemispheric or global reconstruc-
tion (15, 16). Thus, regional studies and recon-
structions of climate change are critically im-
portant when climate impacts are evaluated
(2123). Extremes at regional scales, such as the
hot summer of 2003 in many European areas,
exhibit much larger amplitudes than extremes at
the global scale, and they may thus markedly
affect the local to regional natural environment,
society, and economy, including most vital as-
pects such as water supply and agriculture.
Here we present a new gridded (0.5° 0.5°
resolution) reconstruction of monthly (back to
1659) and seasonal (from 1500 to 1658) tem-
perature fields for European land areas (25°W
to 40°E and 35°Nto70°N) (19). This recon-
struction is based on a comprehensive data set
that includes a large number of homogenized
and quality-checked instrumental data series, a
number of reconstructed sea-ice and tempera-
ture indices derived from documentary records
for earlier centuries, and a few seasonally re-
solved proxy temperature reconstructions from
Greenland ice cores and tree rings from Scan-
dinavia and Siberia (fig. S1 and tables S1 and
S2). We discuss the evolution of European winter,
summer, and annual mean temperatures for more
than 500 years in the context of estimated uncer-
tainties, emphasizing the trends, spatial patterns
for extreme summers and winters, and changes in
both extreme and mean conditions.
Fig. 1A presents the winter [December
through February (DJF)] European surface tem-
perature variations since 1500 (relative to the
1901 to 1995 average) and the 95% confidence
range [2 standard error (SE)] (19). The un-
certainty is larger in the earlier reconstructions.
From the 16th to the beginning of the 18th
century, the two SEs of the filtered wintertime
series are in the order of 1.3°C, and they reduce
to 0.4°C from 1865 onwards. The larger uncer-
tainties in the earlier centuries are mainly due to
a smaller number of uniformly distributed in-
strumental records (none before 1659), but are
also due to fewer proxy series and additional
uncertainties in the documentary data (20, 24
26) and natural proxies (2, 27).
Except for two short periods around 1530
and 1730, European winters were generally
colder than those of the 20th century. The cold-
est multidecadal winter periods were experi-
enced during the late 16th century, during the
last decades of the 17th century, and at the end
of the 19th century (T ⫽⬃0.8°C, where T
is the change from the 1901 to 1995 average).
The winter of 1708/1709 was the coldest in
record (T 3.6°C) and probably related to a
negative North Atlantic Oscillation (NAO) in-
dex (2830). We reconstructed spatial anomaly
patterns of the three individual months and the
average of winter 1708/1709 (Fig. 2A). January
and February contributed most to the overall cold
when temperatures over large parts of Europe and
western Russia were more than 7°C below aver-
age. Except for the northernmost part of our study
area, the reconstruction is reliable. Independent
climate evidence from different European areas
confirms the existence of strong negative temper-
ature anomalies (supporting online text).
We calculated the return period of a
European-wide event such as the coldest winter
of 1708/1709. This calculation is based on fit-
ting a spline function and is sensitive to the
trend over the period 1750 to 2002 and the
assumption of Gaussian distributed residuals
(19). We obtained a return period of 200 to 500
years for winter conditions from 1750 to 1900.
The warming of the 20th century leads to an
increase in the return period, which amounts to
more than 100,000 years at the turn of the 21st
century (fig. S2A). However, the uncertainties of
the estimates are large, and the return periods
should be considered with caution (fig. S2A).
Fig. 2C presents anomaly (1901 to 1995
average subtracted) composites and the corre-
sponding standard deviations (SDs) for the ten
coldest European winters, excluding 1708/
1709. The anomaly composite resembles the
1708/1709 winter. It indicates continental cold
with the largest deviations and highest variabil-
ity over northern and eastern Europe, western
Fig. 1. (A) Winter (DJF), (B)
summer (JJA), and (C) an-
nual averaged-mean Euro-
pean temperature anomaly
(relative to the 1901 to
1995 calibration average)
time series from 1500 to
2003, defined as the aver-
age over the land area
25°W to 40°E and 35°N to
70°N (thin black line). The
values for the period 1500
to 1900 are reconstruc-
tions; data from 1901 to
1998 are derived from
(44). The post-1998 data
stem from (45); Goddard
Institute for Space Studies
(GISS) NASA surface tem-
perature analysis is given
ona1° 1° resolution
(46). Temperature data
from (44) and (45) are very
similar and correlate at
0.98 for each season within
the common period 1901
to 1998 for the chosen
area; they do not indicate
any absolute bias. The
thick red line is a 30-year
Gaussian low-pass filtered
time series. Blue lines show
the 2 SEs of the filtered
reconstructions on either
side of the low-pass fil-
tered values. The red hori-
zontal lines are the 2-SD
line of the period 1901 to
1995. The warmest and
the coldest winters, sum-
mers, and years are denot-
ed in blue and red, respec-
tively. The winter y axis
uses a different scale. Re-
con., reconstructed; CRU,
Climatic Research Unit (44);
TT, temperature; wrt, as
compared to.
R EPORTS
5 MARCH 2004 VOL 303 SCIENCE www.sciencemag.org1500
Russia, and Scandinavia, and positive anoma-
lies over Iceland and parts of Turkey. It shows
the well-known seesaw in winter temperature
between Greenland/Iceland and Europe (31),
associated with large-scale variations in the
atmosphereoceansea ice system.
A strong winter warming trend was ob-
served between 1684 and 1738. The linear trend
for this period amounts to 0.32°C 0.18°C
per decade. (All confidence ranges on trends
are calculated at the 95% confidence level and
are statistically significant.) Such an intense
increase in European winter temperature over a
comparable time period was not observed else-
where in the 500-year record. The spatial trend
map for this 55-year period (Fig. 3) indicates an
increasing warming gradient from southwestern
to northeastern Europe, with maximum values
(0.8°C per decade) over Scandinavia and the Bal-
tic region. A strong trend toward decreased winter
ice severity in the Western Baltic for the same
period has been found (32), thus supporting our
findings with independent climate information.
The large-scale European warming during
this time may have been caused by different
processes. In a stratosphere-resolving general
circulation model (22, 33), decadal-scale conti-
Fig. 2. (A) Monthly (D, J, and F) and seasonal
(DJF) European temperature anomaly (relative
to 1901 to 1995 average) patterns for the
coldest winter (1708/1709). The reconstruc-
tions are based on time series from central
England (United Kingdom, instrumental), Paris
(France, instrumental), Berlin (Germany, in-
strumental) and De Bilt ( The Netherlands,
instrumental). For Switzerland, southern Ger-
many, Hungary, Greece, and Portugal, temper-
ature estimations are available based on high-
resolution documentary evidence. Icelandic
sea ice conditions around the northern, east-
ern, and southern coasts were also used (fig.
S1 and table S2). The Reduction of Error (RE)
values (a measure of common variance) (47)
are plotted on the seasonal winter pattern.
The RE values were calculated from calibration
over the 1901 to 1960 period, with the same
available predictors as for winter 1708/1709,
and were verified over the period 1961 to
1995. [An RE of 1 is a perfect reconstruction,
and RE 0 is better than climatology with
reconstruction skill (19)]. (B) As (A), but for
the hottest summer 2003. The data for 2003
have been provided by (45). (C) Anomaly
(1901 to 1995 average subtracted) composite
and SD of the ten coldest European winters
(excluding winter 1708/1709) and ten hottest summers (excluding summer 2003) over the period 1500 to 2003. Anomalies and SDs are given in °C. White grid cells
indicate missing values. Extremely cold European winters are usually related to negative NAO situations and persistent high pressure systems centered over northern
Europe or Scandinavia or to the westward extension of the Siberian anticyclone connected with continental easterly to northerly flow (48). Hot European summers
are generally connected with a large number of persistent high-pressure systems over the continent, which may result in strong subsidence and/or warm air advection
from the southwest. Dry soils can support the conversion of the surface radiation into heat (at the expense of evapotranspiration).
R EPORTS
www.sciencemag.org SCIENCE VOL 303 5 MARCH 2004 1501
nental winter temperatures before the industrial
era appear to respond differently to solar and
volcanic forcings. Although both enhanced ir-
radiance and large eruptions lead to continental
warming, solar changes affect continental
scales much more strongly, through forcing of
the NAO or Arctic Oscillation (AO) [i.e., en-
hanced (reduced) solar irradiance causes a shift
toward the high (low) index NAO/AO state].
Thus, solar forcing seems to dominate over
volcanic eruptions, which induce a more homo-
geneous hemisphere-wide cooling at decadal
time scales. Increased solar irradiance at the end
of the 17th century and through the first half of
the 18th century might have induced such a
shift toward a high NAO/AO index, which
agrees with independent proxy NAO recon-
structions (28, 29). It is well known that the
NAO exerts a dominant influence on winter-
time temperature over much of Europe, though
the strength of the relationship can change over
time and region (34). We confirm this behavior
for the pre-instrumental period (35). Further-
more, North Atlantic sea surface temperatures
(36) and tropical variability (37) are both rele-
vant for NAO dynamics and might also be
important for explaining the European winter
warming during the late 17th and early 18th
centuries. However, missing ocean data for this
period impede testing of this hypothesis. Solar
irradiance estimates stayed at relatively high
values until the turn of the 19th century, where-
as NAO index reconstructions and European
winter temperatures indicate lower values.
Mechanisms responsible for such European
winter cooling are still under debate.
The linear winter temperature trend for
the 20th century (1901 to 2000) is 0.08°C
0.07°C per decade. The winter 1989/1990
(T ⫽⫹2.4°C) and the decade 1989 to 1998
(T ⫽⫹1.2°C) were the warmest since 1500.
The period 1989 to 1998 was almost two SEs
warmer than the second warmest (non-
overlapping) decade (1733 to 1742, T
0.45°C), thus was very likely (95% confi-
dence level) warmer than any other decade
since 1500. At the multidecadal time scale (30-
year averages), the winters between 1973 and
2002 were likely (85% probability) the warmest
30-year period of the last half-millennium.
Winter 2002/2003, however, was 0.4°C colder
than the 1901 to 1995 average. Recent findings
(38) show that the effect of anthropogenic forc-
ing is detectable on Eurasian winter tempera-
tures over the period 1950 to 1999.
European summer [June through August
(JJA)] temperatures are shown in Fig. 1B. The
two-SE limits decrease from 0.7°C at 1500 to
0.2°C toward the end of the reconstruction period.
A notable increase in reliability occurs in the first
part of the 18th century, when instrumental data
become available. Reconstructed European sum-
mers from 1530 to 1570 were slightly warmer
than the 1901 to 1995 average. A marked feature
in the summer series is the higher temperatures
from 1750 until the second half of the 19th
century, including the second hottest summer of
1757 (T ⫽⫹1.6°C). Importantly, possible inho-
mogeneities in the instrumental data before the
mid-19th century cannot be fully excluded and are
still a matter of discussion. For example, summer
temperature observations from Stockholm and
Uppsala (Sweden) could have been positively bi-
ased by as much as 0.7 to 0.8°C before 1860,
likely because of insufficient radiation protection
of the thermometers (39). Reconstructed Northern
Hemisphere summer temperatures (12) during
this period remained below the 20th-century av-
erage. This underlines the fundamental difference
between late 20th-century warming at the hemi-
spherical scale and preindustrial regional warm
episodes that were as warm or even warmer than
today, but were limited in their geographical ex-
tent and scattered in their timing (15, 40).
From 1757 onwards, there was a summer
cooling trend ( 0.06°C 0.02°C per decade)
until the beginning of the 20th century, with
1902 as the coolest summer in the entire record.
During the 20th century, the instrumental sum-
mer data depict first a warming trend until
1947, followed by a cooling trend until 1977.
Subsequently, an exceptionally strong, unprec-
edented warming is observed (a linear trend of
0.7°C 0.20°C per decade) that featured
very likely the hottest summer decade 1994 to
2003. The European summer temperatures
show other multidecadal periods with compa-
rable, though less strong, warming trends (1731
to 1757, 0.42°C 0.17°C per decade; 1923 to
1947, 0.45°C 0.23°C per decade, respective-
ly). The summer of 2003 exceeded 1901 to
1995 European summer temperatures by
around 2°C (4 SDs). Taking into account the
uncertainties in our reconstructions, it appears
that the summer of 2003 was very likely warm-
er than any other summer back to 1500.
The return period of a European-scale sum-
mer event exceeding 2°C (relative to the 1901
to 1995 average) was calculated with the same
methodology (varying trend over time) as for
the 1708/1709 winter (19). The return period is
more than 5000 years for mid-18th century
summer conditions (fig. S2B). It increases no-
ticeably to millions of years at the turn of the
20th century and decreases to less than 100
years for the most recent summers (19). Schär
et al. (41) found a much higher return period in
their analysis of central European temperature.
The differences might be related to the use of
different methods, such as varying trend over
time versus specified climatology, or of another
base period connected with a different SD, as
well as the different geographical area (regional
versus continental). However, both estimates
contain large uncertainties and should not be
overinterpreted (fig. S2B) (41).
Results from regional climate model simu-
lations (41, 42) (under the Special Report on
Emissions Scenarios A2, transient greenhouse-
gas scenario) suggest that about every second
summer will be as hot or even hotter than 2003
by the end of the 21st century (2071 to 2100).
All individual summer months of 2003 (Fig.
2B) experienced significantly higher-than-
normal temperatures, with maximum values
over western and central Europe. The anomaly
(1901 to 1995 average subtracted) composite of
the ten hottest European summers (excluding
2003) (Fig. 2C) shows a monopole pattern with
the most positive anomalies and highest vari-
ability over northeastern Europe.
The smoothed curve of European annual
mean temperatures (Fig. 1C) clearly points to
cooler conditions throughout the earlier re-
constructed centuries. The 19th century
(T 0.32°C) was the coldest of the last
half-millennium. This agrees with recon-
structions for the Northern Hemisphere (9).
The coldest decadal periods were observed in
the second part of the 19th century, at the end
of the 17th century, and 1600 (T
0.6°C), although with an increasing degree
of uncertainty the earlier in time.
Decadal-scale continental annual temperature
changes during preindustrial times appear to
be driven primarily by solar variability (22,
33), although prolonged periods of volcanism
could have also contributed to European
cooling. Deforestation (6) may also be rele-
vant for lower European annual temperatures
in the late 19th century.
Fig. 3. Winter temperature trends
(°C per decade) from 1684 to
1738. The thick solid lines repre-
sent the 95% and 99% confidence
level (error probability 0.05 and
0.01), respectively, using a Mann-
Kendall trend test. Except for the
Mediterranean area, the warming
trends are statistically significant
over the whole of Europe.
R EPORTS
5 MARCH 2004 VOL 303 SCIENCE www.sciencemag.org1502
The 20th century (1901 to 2000) was the
warmest since 1500. There was a strong warm-
ing trend of 0.08°C 0.03°C per decade
within the 20th century. The last 30 years (1974
to 2003, T ⫽⫹0.43°C) were 0.45°C higher
than the second warmest 30-year periods (1722
to 1751 and 1750 to 1779) of the reconstruc-
tions. When we consider the uncertainties of
earlier periods, the late 20th- and early 21st-
century European warmth at multidecadal (30-
year) scale is very likely unprecedented for
more than the past 500 years. The nine warmest
European years on record have occurred since
1989. The year 1989 (T ⫽⫹1.3°C) and the
decade 1994 to 2003 (T ⫽⫹0.84°C) were
very likely the warmest for more than half
a millennium.
Our 500-year continental-scale surface tem-
peratures provide evidence of current European
climate change. Comparing recent temperature
changes with those of the past and taking into
account reconstruction uncertainties, we show that
the late 20th- and early 21st-century warmth very
likely exceeds that of any time during at least the
past 500 years. The high-resolution reconstruc-
tion also sheds light on the spatial structure of
regional temperature anomalies and extremes
back in time. Furthermore, our temperature es-
timates provide a key test of the General Cir-
culation Models continental and seasonal re-
sponse to different forcings (22, 43).
References and Notes
1. M. E. Mann, R. S. Bradley, M. K. Hughes, Nature 392,
779 (1998).
2. P. D. Jones, T. J. Osborn, K. R. Briffa, Science 292, 662
(2001).
3. K. R. Briffa, T. J. Osborn, Science 295, 2227 (2002).
4. T. J. Crowley, Science 289, 270 (2000).
5. S. Gerber et al., Clim. Dyn. 20, 281 (2003).
6. E. Bauer, M. Claussen, V. Brovkin, A. Huenerbein,
Geophys. Res. Lett. 30, 1276 (2003).
7. R. S. Bradley, P. D. Jones, Holocene 3, 367 (1993).
8. J. T. Overpeck et al., Science 278, 1251 (1997).
9. M. E. Mann, R. S. Bradley, M. K. Hughes, Geophys. Res.
Lett. 26, 759 (1999).
10. K. R. Briffa, Quat. Sci. Rev 19, 87 (2000).
11. K. R. Briffa et al., J. Geophys. Res. 106, 2929 (2001).
12. P. D. Jones, K. R. Briffa, T. P. Barnett, S. F. B. Tett,
Holocene 8, 455 (1998).
13. J. Esper, E. R. Cook, F. H. Schweingruber, Science 295,
2250 (2002).
14. M. E. Mann, S. Rutherford, R. S. Bradley, M. K. Hughes,
F. T. Keimig, J. Geophys. Res. 108, 4203 (2003).
15. M. E. Mann et al., Eos 84, 256 (2003).
16. M. E. Mann et al., Earth Interact. 4-4, 1 (2000).
17. J. Guiot, in European Paleoclimate and Man, Pala¨okli-
maforschung, 7, 93 (1991).
18. D. A. Fisher, Holocene 12, 401 (2002).
19. Materials and methods are available as supporting
material on Science Online.
20. P. D. Jones, K. R. Briffa, T. J. Osborn, J. Geophys. Res.
108, 4588 (2003).
21. Z. W. Kundzewicz, M. L. Parry, in Climate Change
2001: Impacts, Adaptation, and Vulnerability,J.J.
McCarthy et al., Eds. (Cambridge Univ. Press, New
York, 2001), pp. 641– 692.
22. D. T. Shindell, G. A. Schmidt, R. L. Miller, M. E. Mann,
J. Clim. 16, 4094 (2003).
23. C. Pfister, in Kulturelle Konsequenzen der Kleinen
Eiszeit [Cultural Consequences of the Little Ice Age],
W. Behringer, H. Lehmann, C. Pfister, Eds. (Vanden-
hoek & Ruprecht, Go¨ttingen), in press.
24. C. Pfister, Wetternachhersage: 500 Jahre Klimavaria-
tionen und Naturkatastrophen 1496 –1995 (Haupt-
Verlag, Bern, 1999).
25. R. Bra´zdil, C. Pfister, H. Wanner, H. von Storch, J.
Luterbacher, in preperation.
26. A. Pauling, J. Luterbacher, H. Wanner, Geophys. Res.
Lett. 30, 1787 (2003).
27. K. R. Briffa, T. J. Osborn, Science 284, 926 (1999).
28. F. S. Rodrigo, D. Pozo-Vazquez, M. J. Esteban-Parra, Y.
Castro-Diez, J. Geophys. Res. 106, 14805 (2001).
29. E. R. Cook, R. D. D’Arrigo, M. E. Mann, J. Clim. 15,
1754 (2002).
30. J. Luterbacher et al., Atmos. Sci. Lett. 2, 114 (2002).
31. H. van Loon, J. C. Rogers, Mon. Wea. Rev. 106, 296 (1978).
32. G. Koslowski, R. Glaser, Clim. Change 41, 175 (1999).
33. D. T. Shindell, G. A. Schmidt, M. E. Mann, D. Rind,
A. M. Waple, Science 294, 2149 (2001).
34. P. D. Jones, T. J. Osborn, K. R. Briffa, in The North Atlantic
Oscillation: Climatic Significance and Environmental Im-
pact [Geophysical Monograph 134], J. W. Hurrell, Y.
Kushnir, G. Ottersen, M. Visbeck, Eds. (American Geo-
physical Union, Washington, DC, 2003).
35. J. Luterbacher, D. Dietrich, E. Xoplaki, M. Grosjean, H.
Wanner, unpublished data.
36. M. J. Rodwell. D. P. Rowell, C. F. Folland, Nature 398,
320 (1999).
37. J. W. Hurrell, M. P. Hoerling, A. S. Phillips, T. Xu, Clim.
Dyn., in press.
38. F. W. Zwiers, X. Zhang, J. Clim. 16, 793 (2003).
39. A. Moberg, H. Alexandersson, H. Bergstro¨m,P.D.
Jones, Int. J. Climatol. 23, 1495 (2003).
40. R. S. Bradley, M. K. Hughes, H. F. Diaz, Science 302,
404 (2003).
41. C. Scha¨r et al., Nature 427, 332 (2004).
42. N. Nakicenovic et al., Intergovernmental Panel on
Climate Change Special Report on Emissions Scenarios
(Cambridge Univ. Press, Cambridge, UK, 2000).
43. E. Zorita et al., “Simulation of the climate of the last
five centuries” (GKSS Report 2003/12, Geesthacht,
Germany, 2003).
44. M. New, M. Hulme, P. D. Jones, J. Clim. 13, 2217 (2000).
45. J. Hansen et al., J. Geophys. Res. 106, 23947 (2001).
46. Data are available at www.giss.nasa.gov/data/
update/gistemp/.
47. E. R. Cook, K. R. Briffa, P. D. Jones, Int. J. Climatol. 14,
379 (1994).
48. J. Luterbacher et al., Clim. Dyn. 18, 545 (2002).
49. We thank M. E. Mann, P. D. Jones, A. Moberg, F. J.
Gonza´lez-Rouco, T. Jonsson, D. T. Shindell, T. F. Stocker,
J. Esper, C. Pfister, N. Schneider, P. Della-Marta, A.
Pauling, C. Casty, D. Oesch, E. Fischer, T. Rutishauser, P.
Michna, and C. Neuhaus for discussions on various
aspects of this paper; J. Hansen and R. Ruedy from
NASA Goddard Institute for Space Studies for providing
their surface temperature analysis; P. Michna for trend
calculations; P. Della-Marta for English corrections; E.
Fischer for the station network figure; E. Lerch for data
extraction; and many other persons for providing their
instrumental or proxy data. The authors are grateful for
the use of predictor data from various sources. The
Global Climate Data (Version 1, gridded temperature
and precipitation) has been supplied by the Climate
Impacts LINK Project (DEFRA, contract EPG 1/1/154) on
behalf of the Climatic Research Unit, University of East
Anglia. Supported by the Swiss National Science Foun-
dation (NCCR Climate) (J.L.); the Fifth Framework Pro-
gramme of the European Union [Project SOAP (Simu-
lations, Observations and Paleoclimate Data: Climate
Variability over the last 500 Years)] (E.X.); and the
Bundesamt fu¨r Bildung und Wissenschaft under con-
tract 01.0560 (B.B.W.).
Supporting Online Material
www.sciencemag.org/cgi/content/full/303/5663/1499/
DC1
Materials and Methods
SOM Text
References and Notes
Figs. S1 and S2
Tables S1 and S2
20 November 2003; accepted 5 February 2004
Late Miocene Teeth from Middle
Awash, Ethiopia, and Early
Hominid Dental Evolution
Yohannes Haile-Selassie,
1
* Gen Suwa,
2
Tim D. White
3
Late Miocene fossil hominid teeth recovered from Ethiopia’s Middle Awash
are assigned to Ardipithecus kadabba. Their primitive morphology and wear
pattern demonstrate that A. kadabba is distinct from Ardipithecus ramidus.
These fossils suggest that the last common ancestor of apes and humans had
a functionally honing canine–third premolar complex. Comparison with
teeth of Sahelanthropus and Orrorin, the two other named late Miocene
hominid genera, implies that these putative taxa are very similar to A.
kadabba. It is therefore premature to posit extensive late Miocene hominid
diversity on the basis of currently available samples.
The phylogenetic status of the earliest
hominid genera Sahelanthropus, Orrorin,
and Ardipithecus (16 ) and the definition
of the family Hominidae (710) are in de-
bate. By what derived characters should
the hominid (1, 11) clade be recognized?
Bipedality might be an arbiter of hominid
status, but bipedality involves a large
and complex set of anatomical traits and is
not a dichotomous character. Femora at-
tributed to Orrorin tugenensis at 5.8
million years ago (Ma) constitute the ear-
liest postcranial evidence for early hominid
bipedality (2, 12). However, the O. tu-
genensis femora are different from those of
later hominids such as Australopithecus
afarensis (13). Indeed, some question
1
Cleveland Museum of Natural History, 1 Wade Oval
Drive, Cleveland, OH 44106, USA.
2
The University
Museum, The University of Tokyo, Bunkyo-Ku, Hongo,
Tokyo 113-0033, Japan.
3
Department of Integrative
Biology and Laboratory for Human Evolutionary Stud-
ies, Museum of Vertebrate Zoology, University of
California, Berkeley, CA 94720, USA.
*To whom correspondence should be addressed. E-
mail: yhailese@cmnh.org
R EPORTS
www.sciencemag.org SCIENCE VOL 303 5 MARCH 2004 1503
... One of them corresponds to those ofLuterbacher's et al. (2006), which presents the average European temperatures organized by seasons. A second reconstruction, which is also the most recent to date, is that developed byLuterbacher et al. (2004). ...
Article
Full-text available
This article examines the impact of climatic variability on the English Agricultural Revolution using Allen’s Nitrogen Hypothesis. While half of the variation in yields can be attributed to nitrogen-fixing plants, better cultivation, and improved seeds, the remainder can be attributed to changing climatic conditions during the relatively cold period from c. 1645–1715 and the subsequent warmer phase. The study finds that farmers made even greater efforts than observed yields during the colder and more humid climate of the second half of the seventeenth century and the early eighteenth. Conversely, increasing temperatures in the following period had a positive effect on agricultural productivity, indicating that farmers' role during this phase have been overrated.
... The theoretical arguments suggest that in a rural and agriculture dependent Europe, social trust became a key element for farmers to face the difficulties derived from changes in the climatic conditions. In particular, we use the variability of temperatures between years 1500 and 1750, provided by Luterbacher et al. (2004). The correlation with social trust is positive (0.45), significant at the 1% level. ...
Article
Full-text available
Social trust is a deeply-rooted feature of society, whose positive impact on economic performance has been widely documented in many contexts. However, its impact on the non-economic aspects of social progress that characterize advanced societies, such as personal rights, freedom, tolerance and inclusion and access to advanced education is still understudied, especially at the subnational level. As shown by the European Social Progress Index (EU-SPI) 2020, elaborated by the European Commission, the European regions present remarkable disparities in those non-economic aspects. Using the EU-SPI framework, this paper provides fresh evidence on a positive impact of social trust on several features defining advanced social progress. Social trust effects are mainly seen in improved quality of government, education and people's pro-social behaviors. These insights can be useful for the design of future policies that pursue a more equal Europe beyond purely economic indicators, given that regional social trust can condition their success.
... In the PACA region risk indexes were positive most of the time, though there were continuous periods of negative risk (r < 0) from 1887 to1896 and from 1902 to 1926 (Fig. 4), which were long enough to eliminate any incipient outbreak. Furthermore, Europe underwent the largest period of negative summer temperature anomalies due to volcanism forcing in the second quarter of the 19th century, with the highest peak in 1902 41 . The effect of these cool summers is reflected in the lower accumulation of growing degree days in our model when using the summer thermal anomaly instead of the annual thermal anomaly to calibrate the risk index. ...
Preprint
Full-text available
Unlike most grapevine diseases of American origin, the vector-borne bacterium Xylella fastidiosa (Xf) responsible for Pierce's disease (PD) has not yet spread to continental Europe. The reasons for this lack of invasiveness remain unclear. Here, we present phylogenetic, epidemiological and historical evidence to explain how European vineyards escaped Xf. Using Bayesian temporal reconstruction, we show that the export of American grapevines to France as rootstocks to combat phylloxera (~1872-1895) preceded the spread of the Xf grapevine lineage in the US. In the dated tree, the time of the most recent common ancestor places the introduction of Xf into California around 1875, which agrees with the emergence of the main PD outbreak and the onset of its expansion into the southeastern US around 1895. We also show that between 1870 and 1990, climatic conditions in continental Europe were mostly below the threshold for PD epidemics. This lack of spatiotemporal concurrence between factors that could facilitate the establishment of the Xf grapevine lineage would explain the historical absence of PD in continental Europe. However, our model indicates that there has been an inadvertent expansion of risk in southern European vineyards since the 1990s, which is accelerating with global warming. Our temporal approach identifies the biogeographic conditions that have so far prevented PD, and gives continuity to predictions of increased risk in important southern European wine-producing areas under a forthcoming scenario of +2 and +3 degrees C temperature increases.
... During winter, thermal sensations that are caused by cold stress are also affected school children [18] . Europe has experienced warmer climates in the late 20th and early 21st centuries than the previous centuries [19][20][21] . The reduction of thermal energy to the surroundings is called cold stress and under such circumstances, the body's first response is to reduce blood flow through the skin in order to conserve heat. ...
Article
Full-text available
Cold waves, cold nights and warm nights are major threats to human beings during winter due to climate change in different parts of India. The analysis of these has been studied for four major metropolitan cities (Chennai, Mumbai, Kolkata, and Delhi) of India during the period 1985-2020. The authors have used the 90th and 10th percentile threshold to identify the cold nights, warm nights and cold waves during the winter season. The degree of discomfort and cold stress category are identified using the Humidity Index (HD), and the Universal Thermal Climate Index (UTCI). The results indicate that the cold night event in Mumbai is ~0.36% higher than both Kolkata and Chennai cities but it is ~0.42% higher than Delhi. The number of cold wave events in Delhi is 53.5% higher than in Kolkata during the period of the study. It is also observed from the study of UTCI that the possibility of slight cold stress in the Delhi region during cold nights is 59.36% more than in other metropolitan cities. The study indicates that in the winter season, the climate of Delhi is more dynamic but for Kolkata it is relatively less dynamic.
... analysis based on Landsat archives showed a shift in the riparian forest health status near the confluence with the Rhône River. The timing of this change corresponds to the hottest and driest summer recorded in France: the 2003 drought(Black et al., 2004;Luterbacher et al. ...
Preprint
Full-text available
Alluvial forests are fragile and sensitive to drought induced by climate change and exacerbated by altered flow regimes. Our ability to detect and map their sensitivity to drought is therefore crucial to evaluate the effects of climate change and adjust management practices. In such a context, we explore the potential of multi-scale thermal infrared imagery (TIR) to diagnose the sensitivity of alluvial forests to drought events. In summer 2022, we sampled leaves and phloem on Populus nigra trees from two sites with contrasted hydrological connectivity along the Ain River (France) in order to investigate the seasonality of water stress and act as ground truth for airborne TIR images. To map forest sensitivity to drought, we then used a set of TIR data from four existing airborne campaigns and Landsat archives over a larger spatial and temporal extent. Field data showed that stress conditions were reached for both sites during summer but were higher in the site with lower groundwater connectivity, which was also the case for individual tree crown temperature. At the forest plot scale, canopy temperature was linked to forest connectivity for two of four airborne TIR campaigns, with higher values in the more degraded reaches. The data from the Landsat archives at the landscape scale was used to locate the areas of the riparian forest impacted by a historical drought event, and monitor their recovery. TIR data showed promising results to help detect and map tree water stress in riparian environments. However, stress is not detected in all TIR campaigns, demonstrating that in-field ecophysiological measurements are complementary to validate observations and one-shot acquisitions are not enough to diagnose stress. More integrative indicators of drought stress are needed at a seasonal scale, one-shot acquisitions on a given day can inform potential heat disturbance effects but do not really give information on the cumulative effects of heat pulses over the whole vegetative season (ramp-disturbance effect). Landsat data was useful to identify trends but may be less representative of stress due to coarse spatial resolution and potential confounding factors related to changes in successional stages (tree height and density...) at larger temporal scales.
... Widespread continuous meteorological instrumental observations from the Mediterranean, however, are only available for roughly the last 70 years, though a few instrumental series extend back a few centuries in time. 35 Natural archives such as tree rings, lake and marine sediments, speleothems, or corals, along with certain historical observations (e.g. wine-grape and cereal harvesting dates, sugar content of wine, time of grain and grape harvest, snow-cover conditions, freezing of water bodies, cloud and wind observations, lake levels), are proxy records which researchers use to approximate information about pre-instrumental variations in temperature, precipitation and a number of other meteorological variables, including humidity, drought, soil moisture, sea level changes, sea water temperature, water mass circulation and pH. ...
Chapter
Full-text available
Tim Newfield (Georgetown University, USA) and co-authors provide a detailed analysis and review of current academic knowledge on the First Pandemic - the Justinianic Plague including palaeoclimatic and palaeoecological (pollen analysis) considerations.
Article
Full-text available
(1) Introduction: The subject of the present study concerns the analysis of the existence and long time evolution of the solar ~200 yr (de Vries/Suess) and ~2400 yr (Hallstadt) cycles during the recent part of the Wurm ice epoch and the Holocene, as well as their forcing on the regional East European climate during the last two calendar millennia. The results obtained here are compared with those from our previous studies, as well as with the results obtained by other authors and with other types of data. A possible scenario of solar activity changes during the 21st century, as well as different possible mechanisms of solar–climatic relationships, is discussed. (2) Data and methods: Two types of indirect (historical) data series for solar activity were used: (a) the international radiocarbon tree ring series (INTCAL13) for the last 13,900 years; (b) the Schove series of the calendar years of minima and maxima and the magnitudes of 156 quasi 11 yr sunspot Schwabe–Wolf cycles since 296 AD and up to the sunspot cycle with number 24 (SC24) in the Zurich series; (c) manuscript messages about extreme meteorological and climatic events (Danube and Black Sea near-coast water freezing), extreme summer droughts, etc., in Bulgaria and adjacent territories since 296 and up to 1899 AD, when the Bulgarian meteorological dataset was started. A time series analysis and χ2-test were used. (3) Results and analysis: The amplitude modulation of the 200 yr solar cycle by the 2400 yr (Hallstadt) cycle was confirmed. Two groups of extremely cold winters (ECWs) during the last ~1700 years were established. Both groups without exclusion are concentrated near 11 yr sunspot cycle extremes. The number of ECWs near sunspot cycle minima is about 2 times greater than that of ECWs near sunspot cycle maxima. This result is in agreement with our earlier studies for the instrumental epoch since 1899 AD. The driest “spring-summer-early autumn” seasons in Bulgaria and adjacent territories occur near the initial and middle phases of the grand solar minima of the Oort–Dalton type, which relate to the downward phases and minima of the 200 yr Suess cycle. (4) Discussion: The above results confirm the effect of the Sun’s forcing on climate. However, it cannot be explained by the standard hypothesis for total solar irradiation (TSI) variations. That is why another hypothesis is suggested by the author. The mechanism considered by Svensmark for galactic cosmic ray (GCR) forcing on aerosol nuclei was taken into account. However, in the hypothesis suggested here, the forcing of solar X-ray flux changes (including solar flares) on the low ionosphere (the D-layer) and following interactions with the Earth’s lithosphere due to the terrestrial electric current systems play a key role for aerosol nuclei and cloud generation and dynamics during sunspot maxima epochs. The GCR flux maximum absorption layer at heights of 35–40 km replaces the ionosphere D-layer role during the sunspot minima epochs.
Article
Full-text available
Existing global volcanic radiative aerosol forcing estimates portray the period 700 to 1000 as volcanically quiescent, void of major volcanic eruptions. However, this disagrees with proximal Icelandic geological records and regional Greenland ice-core records of sulfate. Here, we use cryptotephra analyses, high-resolution sulfur isotope analyses, and glaciochemical volcanic tracers on an array of Greenland ice cores to characterise volcanic activity and climatically important sulfuric aerosols across the period 700 to 1000. We identify a prolonged episode of volcanic sulfur dioxide emissions (751–940) dominated by Icelandic volcanism, that we term the Icelandic Active Period. This period commences with the Hrafnkatla episode (751–763), which coincided with strong winter cooling anomalies across Europe. This study reveals an important contribution of prolonged volcanic sulfate emissions to the pre-industrial atmospheric aerosol burden, currently not considered in existing forcing estimates, and highlights the need for further research to disentangle their associated climate feedbacks.
Article
Full-text available
A new, well-verified, multiproxy reconstruction of the winter North Atlantic Oscillation (NAO) index is described that can be used to examine the variability of the NAO prior to twentieth century greenhouse forcing. It covers the period A.D. 1400-1979 and successfully verifies against independent estimates of the winter NAO index from European instrumental and noninstrumental data as far back as 1500. The best validation occurs at interannual timescales and the weakest at multidecadal periods. This result is a significant improvement over previous proxy-based estimates, which often failed to verify prior to 1850, and is related to the use of an extended reconstruction model calibration period that reduced an apparent bias in selected proxies associated with the impact of anomalous twentieth century winter NAO variability on climate teleconnections over North Atlantic sector land areas. Although twentieth century NAO variability is somewhat unusual, comparable periods of persistent positive-phase NAO are reconstructed to have occurred in the past, especially before 1650.
Article
Full-text available
We derive an optimal Northern Hemisphere mean surface temperature reconstruction from terrestrial borehole temperature profiles spanning the past five centuries. The pattern of borehole ground surface temperature (GST) reconstructions displays prominent discrepancies with instrumental surface air temperature (SAT) estimates during the 20th century, suggesting the presence of a considerable amount of noise and/or bias in any underlying spatial SAT signal. The vast majority of variance in the borehole dataset is efficiently retained by its two leading eigenvectors. A sizable share of the variance in the first eigenvector appears to be associated with non-SAT related bias in the borehole data. A weak but detectable SAT signal appears to be described by a combination of the first two eigenvectors. Exploiting this eigendecomposition, application of optimal signal estimation methods yields a hemispheric borehole SAT reconstruction that is largely consistent with instrumental data available in past centuries, and is indistinguishable in its major features from several published long-term temperature estimates based on both climate proxy data and model simulations.
Article
Full-text available
Using an optimal detection technique, the extent to which the combined effect of changes in greenhouse gases and sulfate aerosols (GS) may be detected in observed surface temperatures is assessed in six spatial domains decreasing in size from the globe to Eurasia and North America, separately. The GS signal is detected in the annual mean near-surface temperatures of the past 50 yr in all domains. It is also detected in some seasonal mean temperatures of the past 50 yr, with detection in more seasons in larger domains.
Article
Full-text available
Palaeoclimatology provides our only means of assessing climatic variations before the beginning of instrumental records. The various proxy variables used, however, have a number of limitations which must be adequately addressed and understood. Besides their obvious spatial and seasonal limitations, different proxies are also potentially limited in their ability to represent climatic variations over a range of different timescales. Simple correlations with instrumental data over the period since ad 1881 give some guide to which are the better proxies, indicating that coral- and ice-core-based reconstructions are poorer than tree-ring and historical ones. However, the quality of many proxy time series can deteriorate during earlier times. Suggestions are made for assessing proxy quality over longer periods than the last century by intercomparing neighbouring proxies and, by comparisons with less temporally resolved proxies such as borehole temperatures. We have averaged 17 temperature reconstructions (representing various seasons of the year), all extending back at least to the mid-seventeenth century, to form two annually resolved hemispheric series (NH10 and SH7). Over the 1901–91 period, NH10 has 36% variance in common with average NH summer (June to August) temperatures and 70% on decadal timescales. SH7 has 16% variance in common with average SH summer (December to February) temperatures and 49% on decadal timescales, markedly poorer than the reconstructed NH series. The coldest year of the millennium over the NH is ad 1601, the coldest decade 1691–1700 and the seventeenth is the coldest century. A Principal Components Analysis (PCA) is performed on yearly values for the 17 reconstructions over the period ad 1660–1970. The correlation between PC1 and NH10 is 0.92, even though PC1 explains only 13.6% of the total variance of all 17 series. Similar PCA is performed on thousand-year-long General Circulation Model (GCM) data from the Geophysical Fluid Dynamics Laboratory (GFDL) and the Hadley Centre (HADCM2), sampling these for the same locations and seasons as the proxy data. For GFDL, the correlation between its PC1 and its NH10 is 0.89, while for HADCM2 the PCs group markedly differently. Cross-spectral analyses are performed on the proxy data and the GFDL model data at two different frequency bands (0.02 and 0.03 cycles per year). Both analyses suggest that there is no large-scale coherency in the series on these timescales. This implies that if the proxy data are meaningful, it should be relatively straightforward to detect a coherent near-global anthropogenic signal in surface temperature data.
Article
Full-text available
Climatic changes resulting from greenhouse gases will be superimposed on natural climatic variations. High-resolution proxy records of past climate can be used to extend our perspective on regional and hemispheric changes of climate back in time by several hundred years. Using historical, tree-ring and ice core data, we examine climatic variations during the period commonly called the 'Little Ice Age'. The coldest conditions of the last 560 years were between AD 1570 and 1730, and in the nineteenth century. Unusually warm conditions have prevailed since the 1920s, probably related to a relative absence of major explosive volcanic eruptions and higher levels of greenhouse gases.
Article
Full-text available
For the purpose of climate signal detection, we introduce a method for identifying significant episodes of large-scale oscillatory variability. The method is based on a multivariate wavelet algorithm that identifies coherent patterns of variation simultaneously within particular ranges of time and periodicity (or frequency) that may vary regionally in the timing and amplitude of the particular temperature oscillation. By using this methodology, an analysis is performed of the instrumental record of global temperatures spanning the past 140 years. The duration of an "episode" is chosen to correspond to 3-5 cycles at a specified oscillation period, which is useful for detecting signals associated with the global El Niño/Southern Oscillation (ENSO) phenomenon. To confirm the robustness of signals detected in the earliest, sparse data (only 111 5° longitude by 5° latitude grid points are available back to 1854), we performed multiple analyses overlapping in time, using increasingly dense subsets of the full (1570 grid point) temperature data. In every case, significant interannual episodes are centered in the 3-7 year period range corresponding to the conventional band of ENSO-related variance and describe intervals of quasi-oscillatory variability of decadal-scale duration. These episodes consist of a sequence of one or two warm and cold events with sea surface temperature fluctuations in the eastern tropical Pacific of amplitude ±0.6°-1.1°C. Each episode includes one or more historically prominent El Niño events. The signals are characterized as significant, however, by virtue of their global-scale pattern of temperature variations as well as their oscillatory pattern in time. The 1920-1940 interval of increasing global temperatures was bracketed by oscillatory
Article
This Working Group II volume brings us completely up-to-date on the vulnerability of socio-economic and natural systems to climate change.
Article
Palaeotemperature sensitive series from tree-rings, ice cores, corals and documentary sources are combined to producepalaeo summer-temperaturereconstructions (AD 1970-1761)and geographical eigenvector (EOF) maps in both hemispheres. They are compared favourably to those of existing summer-temperature average series. There are 51 palaeoseries in the Northern Hemisphere (mostly north of 40° N) and 16 series in the Southern Hemisphere. The statistics and significance of the palaeoreconstructions are examined by: (1) finding the correlation coefficient (palaeo to measured) as a function of the number and geographical distribution of the palaeo series; (2) developing and running a multiproxy model that generates pseudo series containing a signal and the same types and amounts of noise found in the various real palaeo series. The model reproduces the measured correlation coefficients and the eigenvector's (EOF) explained variances as functions of the number of sites. About 77% of the signal variance can be recovered with 51 well-distributed palaeo series and about 90% with greater than 100 series. The 1st eigenvector (EOF l) component maps in the Northern and Southern Hemispheres are fundamentally different, with the Northern having much less longitudinal variation than the Southern. This means (statistically) that simple hemispheric averages of summer temperature have more meaning in the Northern Hemisphere than in the Southern. Even though the various palaeoseries have different spectral biases, noise types and amounts, they were all used together and the common signal extracted. The results of this work strongly suggest that the multiproxy method is valid and that with enough sites most of the summer signal and geographical pattern can be extracted.
Article
We describe new reconstructions of northern extratropical summer temperatures for nine subcontinental-scale regions and a composite series representing quasi "Northern Hemisphere" temperature change over the last 600 years. These series are based on tree ring density data that have been processed using a novel statistical technique (age band decomposition) designed to preserve greater long-timescale variability than in previous analyses. We provide time-dependent and timescale-dependent uncertainty estimates for all of the reconstructions. The new regional estimates are generally cooler in almost all precalibration periods, compared to estimates obtained using earlier processing methods, particularly during the 17th century. One exception is the reconstruction for northern Siberia, where 15th century summers are now estimated to be warmer than those observed in the 20th century. In producing a new Northern Hemisphere series we demonstrate the sensitivity of the results to the methodology used once the number of regions with data, and the reliability of each regional series, begins to decrease. We compare our new hemisphere series to other published large-regional temperature histories, most of which lie within the 1σ confidence band of our estimates over most of the last 600 years. The 20th century is clearly shown by all of the palaeoseries composites to be the warmest during this period.
Article
The climate response to variability in volcanic aerosols and solar irradiance, the primary forcings during the preindustrial era, is examined in a stratosphere-resolving general circulation model. The best agreement with historical and proxy data is obtained using both forcings, each of which has a significant effect on global mean temperatures. However, their regional climate impacts in the Northern Hemisphere are quite different. While the short-term continental winter warming response to volcanism is well known, it is shown that due to opposing dynamical and radiative effects, the long-term (decadal mean) regional response is not significant compared to unforced variability for either the winter or the annual average. In contrast, the long-term regional response to solar forcing greatly exceeds unforced variability for both time averages, as the dynamical and radiative effects reinforce one another, and produces climate anomalies similar to those seen during the Little Ice Age. Thus, long-term regional changes during the preindustrial appear to have been dominated by solar forcing.