Seasonal and Regional Variation of Pan-Arctic Surface Air Temperature
Over the Instrumental Record
James E. Overland
1
Michael C. Spillane
2
Donald B. Percival
3
Muyin Wang
2
Harold O. Mofjeld
1
NOAA/Pacific Marine Environmental Laboratory
1
7600 Sand Point Way NE
Seattle, WA 98115-6349
Joint Institute for the Study of the Atmosphere and Oceans
2
Box 354235, University of Washington
Seattle, WA 98195-4235
Applied Physics Laboratory
3
Box 355640
University of Washington
Seattle, WA 98195-5640
Revision submitted to Journal of Climate
6 February 2004
Contribution 2546 from NOAA/Pacific Marine Environmental Laboratory
1
Abstract
Instrumental surface air temperature (SAT) records beginning in the late 1800s from 59
Arctic stations north of 64°N show monthly mean anomalies of several degrees and large spatial
teleconnectivity, yet there are systematic seasonal and regional differences. Analyses are based
on time/longitude plots of SAT anomalies and Principal Component Analysis (PCA). Using
monthly station data rather than gridded fields for this analysis highlights the importance of
considering record length in calculating reliable Arctic change estimates; for example, we
contrast PCA performed on 11 stations beginning in 1886, 20 stations beginning in 1912, and 45
stations beginning in 1936. While often there is a well-known interdecadal negative covariability
in winter between northern Europe and Baffin Bay, long-term changes in the remainder of the
Arctic are most evident in spring, with cool temperature anomalies before 1920 and Arctic-wide
warm temperatures in the 1990s. Summer anomalies are generally weaker than spring or winter
but tend to mirror spring conditions before 1920 and in recent decades. Temperature advection in
the trough-ridge structure in the positive phase of the Arctic Oscillation (AO) in the North
Atlantic establishes wintertime temperature anomalies in adjacent regions, while the
zonal/annular nature of the AO in the remainder of the Arctic must break down in spring to
promote meridional temperature advection. There were regional/decadal warm events during
winter and spring in the 1930s to 1950s, but meteorological analysis suggests that these SAT
anomalies are the result of intrinsic variability in region flow patterns. These mid-century events
contrast with the recent Arctic-wide AO influence in the 1990s. The preponderance of evidence
supports the conclusion that warm SAT anomalies in spring for the recent decade are unique in
the instrumental record, both in having the greatest longitudinal extent and in their associated
patterns of warm air advection.
2
1. Introduction
In this paper we further the analysis of Arctic variability by reexamination of the surface
air temperature (SAT) from major weather observation stations with long record lengths. We
focus on changes in each season and in different regions of the Arctic, rather than concentrating
on annual zonal averages, because extensive averaging can often obscure the underlying physics.
Our analysis also provides a reevaluation of the instrumental temperature record in that we base
our methodology on station data rather than gridded analyses as in most earlier studies. This
approach avoids the possible introduction of artifacts due to gridding, which is particularly
important in understanding the results from Principal Component Analysis (PCA).
The most difficult issue in retrospective analyses of the Arctic is the range of starting
dates in the instrumental record. Although data coverage is far from complete, there is
considerable information from land areas of the Arctic dating from the early 20th century.
Przybylak (2000) notes that there is good Arctic coverage since the 1950s. Due to the strong
spatial correlation within Arctic subregions (Przybylak 2003) however, there is adequate
coverage since the 1930s. Unfortunately, stations beginning in the mid-1930s do not fully resolve
the mid-century warming episodes. Records from the 1880s are limited to three subregions, west
Greenland, Iceland, and northern Europe. Representative hemispheric spatial coverage (20
stations) is available beginning in 1912 when Svalbard came on line, and we focus primarily on
this period.
The fourth International Polar Year is scheduled for 2007 and will emphasize unresolved
issues of polar influence on climate variability. It is fitting to review the data that began with the
First Polar Year (1882–1883), which marked a transition from exploration to scientific study in
the Arctic. It is also fitting to update the analyses of the many authors in the 1920s–1940s who
noted the warm temperature anomalies of the period (Ahlmann 1948) and pioneered the concept
of high-latitude climate variability, in contrast to the prevailing uniformitarianism.
Recent studies show considerable change in the Arctic over the previous three decades in
both physical and biological indicators (Serreze et al. 2000; Overland et al. 2004). These
3
indicators suggest a shift in atmospheric patterns such as the Arctic Oscillation (AO) and related
stratospheric cooling around 1989, while subarctic records such as permafrost temperatures show
linear trends from the 1970s. It is important to put Arctic change of the past 30 years in the
context of the period from the early 1800s to present, noting the warm temperature anomalies in
the mid 20th century as the end of a long period of rising temperatures. Recent analyses of the
mid-century warming suggest that internal atmospheric variability is important to its explanation
and that the regional dynamics were different compared to recent decades (Hanssen-Bauer and
Førland 1998; Bengtsson et al. 2003; Johannessen et al. 2003; Dickson, personal communication,
2003). As we shall show, it is also important to note the large seasonal and regional differences
in understanding temperature change on decadal time scales throughout the 20th century.
Two recent studies call into question whether changes in the recent period (1990s) are
unique compared to longitudinally averaged, historical temperature data; both papers note the
warm events in 1920–1950. Przybylak (2002) states that while 1991–2000 is the warmest decade
of the second half of the 20th century, “the question remains whether the tendency will continue
and whether the first decades of the 21st century will exceed the levels of the 1930s and 1940s.”
Polyakov et al. (2002) concludes that Arctic air-temperature trends during the 20th century do
not support the predicted polar amplification of global warming, and propose a 50-year Low
Frequency Oscillation (LFO) in Arctic temperatures.
In contrast, modeling studies for the IPCC report (Stott et al. 2001) and proxy
temperature records (Crowley 2000; Briffa and Osborn 2002) make the case for recent warming
relative to the previous two centuries based on external forcing driven by solar variations,
aerosols from volcanoes, and carbon dioxide. In particular, the cool period in the first half of the
1800s can in part be associated with major aerosol production, while the warm period in the first
half of the 20th century had almost no volcanic production (Robock 2000). Warming in the last
2
half of the 20th century is presented as an argument for the uniqueness of recent CO increases in
overcoming the increase in volcanic influence in the last 40 years. Thus we have two competing
4
visions for the future at high northern latitudes: a 50-year cycle going to colder temperatures or
continued warming.
In investigating this climatic issue it is important to resolve methodological issues. The
usual formulation of PCA requires a complete temporal/spatial data matrix. Previous analyses
beginning in 1881 and 1892 (Kelly et al. 1982; Semenov and Bengtsson 2003) rely on gridded
fields where much of the data matrix for the early years is completed by a fill-in rule. We avoid
the need for gridding by conducting three PCAs with time series limited to those stations
beginning in or before 1886, 1912, and 1936. This approach requires that we fill only a few
temporal data gaps. We also limit our PCA to north of 64°N, remaining Arctic centric, while
earlier studies included considerably more stations southward to 60°N and 40°N.
The next section discusses the available data series. This is followed by a visual and
semiquantitative (PCA) inspection and interpretation of the nearly complete instrumental record
of Arctic SAT.
2. Data sources, preparation, and analysis methods for monthly SAT data
a. Sources of station records
The long-term records of monthly mean SAT are based primarily on the Global Historical
Climatology Network (GHCN) dataset version-2 (http://www.ncdc.noaa.gov/cgi-bin/
res40.pl?page=ghcn.html). The data are organized by World Meteorological Organization
(WMO) location index, with records from adjacent locations supplementing that from the
primary site. A secondary source of monthly mean data is the World Monthly Surface
Climatology (WMSC) dataset (http://dss.ucar.edu/catalogs/ranges/range560.html). The selected
time series were cross checked with the data set from Polyakov et al. (2002).
b. Time series selection
It is important to define a southern border of the Arctic for climatological study. There
are few stations north of 80°N, but a major increase in the number of stations as one extends the
5
southern domain south from 65°N to 60°N. Many authors use a fixed latitude such as 60°N
(Walsh 1977) or 62°N (Polyakov et al. 2002). Climatological/botanical limits include the tundra
line (<10° maximum mean monthly temperature) or a more comprehensive multiseasonal,
multimeteorological element approach (Przybylak 2000). The definition is more complex if one
considers the southern limits of hydrologic river basins, which flow into the Arctic. There is no
firm selection criteria.
For our purpose, we propose a practical southern limit of 64°N. Adding the many stations
south of this latitude bias results to the subarctic region. On the other hand, 64°N provides a
reasonable climatological limit for stations that lie north of the Arctic front in winter and provide
sufficient geographic and temporal station coverage. For example, we include Fairbanks, Alaska
but not Anchorage. We also include some interior stations in northern Canada and Russia that
have long instrumental records. In maps of recent temperature trends (Chapman and Walsh 1993)
there are north-south orientations to anomalies, which suggest these relatively lower latitude
stations can be representative of higher latitudes. Another concern is the Scandinavian peninsula,
which is subject to strong warm air advection. Przybylak (2000) excludes this region while
Polyakov et al. (2002) includes stations south to Thorshaven and Aberdeen. Our criteria includes
northern Scandinavian stations which are often subject to Arctic air masses; we will show later
that these stations correlate well with anomalies at Arctic stations further east. They are also
important as they often form a correlation dipole (North Atlantic Oscillation—NAO) with Arctic
stations in the Baffin Bay region.
To minimize data gaps and maximize the number of useable long time series, data were
combined as follows:
stations were limited to north of 64°N in the GHCN database,
each record was initially populated with data from the primary WMO location,
data gaps were successively filled using information from the supplementary adjacent
location, if available, in the order they appear in the GHCN Version-2 file,
6
updates, and some insertions in gaps, were made based on the WMSC (ds570.0)
dataset,
time series, for each station and month, were quality-checked and a few spurious
values were eliminated, and
stations which begin after 1936 were not used, with a few exceptions where regional
coverage is sparse, such as the Canadian Arctic.
The zonal coverage of the identified stations is not uniform. In data rich areas, such as
Scandinavia and the western Russian Federation, some series were removed; the stations retained
were those deemed optimal in terms of duration of coverage, continuity of data, and
representativeness. The result is a set of 59 stations which will be used for visual inspection.
They are plotted in Fig. 1 and listed in Table 1 by longitude. An overall measure of
completeness for each station is given by the percent of monthly observations present in the
available period for that location.
Uniform station density and continuity are particularly important for the PCA where each
station provides a weight function. For this reason we further removed several adjacent stations
or those with short record before proceeding with the PCA as noted in Figure 1 by smaller station
numbers.
c. Methods
For visual presentation of time series and principal components, we apply a five-year
running average to suppress interannual fluctuations. We justify the five-year smoothing as an
approach for investigating century-long decadal change. Decadal records often shift in response
to changes in the frequency of extreme events. For example, the number of winters with cold
stratospheric temperature anomalies increased in the 1990s relative to the 1980s. Changes in the
frequency of volcanic events can affect SAT on decadal time scales. There are natural changes in
storminess from year-to-year. There is also the potential for a high-latitude influence from the
7
quasi-biennial-oscillation. Thus decadal scales appear to be appropriate to address climate
variability over the instrumental record.
We are interested in contrasting monthly/seasonal differences in SAT anomaly patterns.
Combining Arctic data into conventional seasonal three-month averages obscures the patterns we
seek to understand. Thus we group monthly data into quasi-seasonal series based on their
similarity. We define Arctic “winter” as December–January, “spring” as April, “summer” as
July–August and “fall” as October. Other months weakly resemble these patterns or can be
considered transitional. We discuss the consequences of this approach in section 3c.
We conduct a PCA for comparison with visual inspection of the time series. Two
important issues are gap filling of the data and the variable length of the records. With regard to
the few remaining data gaps, we use both a Monte Carlo approach, where missing data are filled
by sampling from a statistical model and repeating the PCA multiple times, and an estimation
procedure for the PCs (Davis 1976). Both approaches gave similar results and indicate that the
PCAs are not materially influenced by the missing temporal data. The decreasing number of
stations in the early record is more problematic. To keep as close to the observed data as
possible, we perform three separate PCAs on subsets of the Arctic stations based on length of
record: 11 stations beginning in 1886 which are from Greenland and Scandinavia, 20 stations
beginning in 1912, and 45 stations beginning in 1936. These data are available at
www.unaami.noaa.gov.
3. Visual inspection and PCA of instrumental temperature time series
As discussed below, visual inspection suggests strong spatial correlation of temperature
anomalies within different segments of the Arctic at decadal scales. We define six sectors in the
text corresponding to northern Europe, Siberia, Beringia, northwestern Canada, Baffin Bay, and
east Greenland; these regions are located in Fig. 1. The northern Europe sector extends eastward
from Scandinavia to Archangelsk and includes the islands of the northeastern Atlantic. Beringia
represents the region on both sides of Bering Strait. The inland location of Fairbanks groups
8
more strongly with northwestern Canada than with Beringia. The Baffin Bay region consists of
land stations from both western Greenland and northeastern Canada. The east Greenland sector
includes Iceland. These locations are approximately the same climatic regions proposed by
Przybylak (2003), except for his combining east Greenland with northern Europe, which from
visual inspection of the station time series is not unreasonable. The longest records are from
northern Europe, Baffin Bay, and eastern Greenland; there are several sites in other regions
which begin before 1900.
a. Winter
We begin the analysis by inspecting time/longitude plots of temperature anomalies
relative to 1961–1990 means for December–January (Fig. 2). The time/longitude plots for
individual months are available at www.unaami.noaa.gov. The northern Europe sector (Stns.
1–9) shows evidence of an interdecadal signal with alternating cold and warm anomaly periods
throughout the record back to the 1860s; there are two periods with long warm anomalies:
1890–1910 and 1920–1938. The E. Greenland sector (Stns. 54–59) often follows the northern
Europe sector, except that it lacks the cooling period around 1940 and a warm period in the
1970s.
In Siberia (Stns. 10–26) generally cold temperature anomalies occurred before 1920
based on Stns. 16, 19, and 25, followed by a warm period in the 1920s through the 1950s.
Maximum warm anomalies in the late 1930s and early 1940s occur several years later than the
mid-1930s maximum in northern Europe. After the 1970s the tendency for Siberia is to be out of
phase with northern Europe, with a warm period in the 1980s and a cool period in the late 1990s
similar to 1890–1910. Far Eastern Siberia and North America generally stayed cool until the mid
1970s, with Beringia (Stns. 27–36) turning cold again starting in the late 1980s and NW Canada
(Stns. 37–44) remaining warm.
The Baffin Bay region (Stns. 45–53) was especially cold during 1865–1910 when Europe
was warm. It was mostly in phase with Europe from the late 1910s to 1948, with a cold followed
9
by a warm period, but again shows an out-of-phase relation for three cold and two warm events
from 1950 to 1995. In the late 1990s Baffin Bay and northern Europe are both warm.
Figure 3 (top) shows the stations used for the three PCAs. Different symbols denote those
stations which begin in 1886 (11 stations), 1912 (20 stations), and 1936 (45 stations). Note that
the later analyses include the stations from the earlier period. The first pattern (A) for winter,
December/January, is represented by the first EOF modes for all three record periods, plotted on
a longitudinal axis and accounting for at least 30% of the interannual variance. All three EOFs
(station weights) show similar station contributions, although the EOF for the series beginning in
1912 shows an enhanced contribution from the Beringia region. In agreement with Fig. 2, pattern
(A) suggests the well-known out-of-phase behavior in winter (NAO or seesaw) between Baffin
Bay and northern Europe (van Loon and Rogers 1978). The first principal component time series
in winter for the three record lengths (Fig. 3, bottom) has a strong interdecadal signal.
The increase in European winter temperatures in the 1920s was associated with increased
warm air advection into Europe; at the same time Baffin Bay was cold, suggesting an NAO
connection (Rogers 1985, Fu et al. 1999). Fu et al. (1999) notes considerable strengthening and
northward movement of the North Atlantic high pressure region during this period, suggesting
midlatitude processes. Temperatures over the sea ice in the central Arctic during the 1980s and
1990s also have this wintertime seesaw pattern (Rigor et al. 2000); there was a warming in the
European sector of the Arctic but a cooling trend in the remainder of the Arctic. During the
positive phase of the AO in the 1990s tropospheric/stratospheric coupling is considered
important in maintaining this pattern (Newman et al. 2001; Moritz et al. 2002; Ambaum and
Hoskins 2002).
An example of the NAO-seasaw (Fig. 4, left panel) is shown on a polar stereographic
projection for winter 1933 SAT anomalies based on the Climate Research Unit data set TS2.0
(New et al. 1999, 2000; http://dss.ucar.edu/datasets/dso10.1) together with sea-level pressure
(SLP) anomalies (Trenberth and Paolino 1980). Northern Europe and Baffin Bay temperature
anomalies are out of phase and the warm anomalies concentrated over Scandinavia. The
10
corresponding SLP anomaly field indicates anomalous southwesterly flow of warm marine air
toward northern Europe, and anomalous northerly flow over Baffin Bay fills the region with cold
Arctic air. The remainder of the Arctic was generally cold in 1933, even though Arctic-wide
decadal winter and annual temperature anomalies were positive (Polyakov et al. 2002; Semenov
and Bengtsson 2003).
The second winter pattern (B) is based on EOF mode 2 for the 1886 record and mode 3
for the 1912 record (Table 2), and shows an in-phase behavior for Baffin Bay and northern
Europe (Fig. 5). The PCs (Fig. 5, bottom) shows positive values from 1925 through 1955, which
is evident in the time/space plot of Fig. 2. In contrast to the 1920s, the winter temperature
anomalies in the late 1930s has northern Europe and Baffin Bay in phase.
The winter of 1936 (Fig. 4, center panel) corresponds to Pattern B. The anomalous low
pressure anomaly is now limited to the northeastern Atlantic region, and a hemispheric
wavenumber 2 pattern is suggested. In contrast to the winter of 1933, there is an anomalous high
located over Greenland. Warm anomalies over the Baffin Bay region are produced by the warm
temperature advection from lower latitudes. The warm anomalies over northern Europe are
shifted to western Russia with strong anomalous southerly flow. The winds for pattern B are
more meridional, while pattern A is more zonal.
In support of pattern B, Hanssen-Bauer and Førland (1998) note that the continued
warming at Svalbard (Stn. 2) in the 1930s was not associated with NAO type warm air advection
processes as in the 1990s. Skeie (2000) and Bengtsson et al. (2003) relate the warming in the
1930s to a high Arctic mode of internal variability, separate from the more subarctic NAO
influence, and make a case that sea ice variability in the Barents/Kara Sea contributes a positive
feedback to maintain the warm temperature anomalies. However, the Baffin Bay area also
contributes to pattern B, so changes in atmospheric circulation on larger scales are also indicated.
Note that the PCA for the short record beginning in 1936 does not contribute toward
Pattern B; this suggests caution when establishing EOFs based on recent data and projecting
11
them onto earlier data records. PC 2 in winter for the short record (not shown) had an Arctic-
wide in phase behavior not present in the longer analyses.
The third winter pattern (C) (Fig. 6) is represented by EOF mode 2 for the analysis
beginning in 1912 and EOF mode 3 for that beginning in 1936 (Table 2). These EOFs show an
in-phase behavior from Siberia eastward through northwestern Canada. The nearly pan-Arctic in-
phase behavior, excluding Baffin Bay and northern Europe, point to a short simultaneous cold
event in the PC series (Fig. 6, bottom) in the late 1910s and the warm Beringia and Siberia event
in the 1980s. Although a regression fit to the 1936 PC record would lead to a positive slope, the
analysis of Rogers and Mosley-Thompson (1995) suggests that the mild Siberian winters of the
1980s, also seen in Fig. 2, were associated with increased storminess in northwestern Siberia
while NAO appeared to have had little influence. Figure 4 (right panel) shows the SAT
anomalies and the SLP anomalies composites for the winters of 1983 and 1984, when there is
extensive warming in Siberia and Beringia and cooling in Baffin Bay. These years are similar to
our Pattern C. The SLP field shows that an anomalous low center is strong and extends
northeastward toward the Barents Sea, with a major influence of anomalous southerly winds over
eastern Siberia.
In summary based on visual inspection and PCA, the northern European sector shows a
interdecadal signature in winter (December–January) throughout the instrumental record with
extended warm temperatures in the mid 1930s. Siberia had warm anomalies from the 1930s
through the late 1940s, generally in phase but occurring somewhat later than northern Europe. A
major event was the warm anomalies across Siberia and western North America (Stns. 7–46) in
the 1980s; subsequently this event has reversed for all regions except northeastern Canada. Our
decadal analysis reinforces the point made by other authors that care must be taken in selecting
intervals for calculating linear trends. For example, from examining Fig. 2, Siberia and Alaska
(Stns. 24–37) would show a positive trend over the previous 40 years in a regression analysis,
even though the main feature was a single decadal warming episode in the 1980s that was
followed by cool anomalies.
12
For comparison with the 20th century Arctic records, there are instrumental records
which begin in the 1700s from the area just south of our region in northern Europe, e.g.,
Stockholm (Moberg et al. 2002). Figure 7 shows a multi-resolution analysis (MRA) based upon
the maximal overlap discrete wavelet transform and the Haar wavelet (Percival and Mofjeld
1997). The Haar wavelet was chosen because it sharply resolves events in the time domain. This
MRA is an additive decomposition of the December–January temperature record for Stockholm
in terms of a 64-year interval (S5) and bands of 2–4, 4–8, 8–16, 16–32, and 32–64 years (D 1 to
D5). The >64 year curve shows a long-term warming trend over the previous two centuries. The
changes in 16–32 year band (D4) show a strong interdecadal signal in the 20th century that
reflects a NAO signature, indicating the predominance of N. Atlantic versus polar Arctic air
masses. In the 19th century, however, temperatures were colder and the impact of the NAO
appears diminished. This shift is also supported by a “regime change” in Baltic Sea ice coverage
in the late 1800s (Omstedt and Chen 2001).
b. Spring
Spring as represented by April surface temperature anomalies (Fig. 8) is the time when
sunlight returns to the Arctic and the winter stratospheric polar vortex weakens and then breaks
down. The most striking feature in the figure is the longitudinal bandedness of the anomalies;
this bandedness cannot be attributed to longitudinal smoothing since we have applied none. The
bandedness is shown in the large warm anomalies since the late 1980s, with particularly strong
anomalies in Siberia, Beringia, and NW Canada (Stns. 11–50). Except for a short period in the
mid 1970s, the Arctic was cool in spring from the 1960s to the late 1980s. There are previous
isolated warm anomalies in the late 1940s/early 1950s, covering Siberia, Beringia, NW Canada,
and Baffin Bay; these regions match the magnitude of the warmest anomalies in the 1990s.
The northern Europe sector (Stns. 1–9) does not contain strong anomalies during spring
of either sign over the length of the record. However, we do note cool anomalies in spring in the
1920/1930s in contrast to the warm wintertime anomalies. On the other hand, Baffin Bay
13
(Stns. 45–53) was warm in both winter and spring during the 1930s. The first PCA pattern (A)
for April (Fig. 9) is represented by the EOF mode 1 for all three records beginning in 1886, 1912,
and 1936. It shows an in phase behavior for the Arctic, but with little contribution from the
northern Europe, E. Greenland, or Baffin Bay sectors. The PC time series for these three
estimates (Fig. 9, bottom panel) show high values in the 1940s and early 1950s and also over
much of the 1990s, consistent with Fig. 8.
The recent warming is of particular interest. Rigor et al. (2000) found that during spring
almost all the central Arctic shows significant warming trends over the previous 20 years. The
warming in Alaska relates to changes in the frequency of southerly warm air advection events
during the breakdown of the polar vortex (Overland et al. 2002). The importance of this spring
pattern A in contrast to the winter pattern A is highlighted by Rogers and McHugh (2002), who
note the similarity of the AO and NAO patterns in winter, but a separation of an Arctic-centric
AO pattern from a more N. Atlantic-oriented NAO pattern during spring.
The second pattern (B) for spring (Fig. 10), supported by all three EOF mode 2s, suggests
a positive-negative shape with North America out-of-phase with Eurasia. The PC time series of
these EOFs (Fig. 10, bottom) point to a major warm event in Beringia and Canada from the mid
1920s until 1940 that can be seen in Fig. 8.
While there has been considerable discussion in the literature about the mid-century
warming in winter during the 1930s, little attention has been given to the warm periods in spring
during the early 1940s and 1950s. To compare the meteorological conditions in recent springs to
these earlier periods, Fig. 11 (right) shows composite plots of the SAT anomalies for April and
SLP for March–April for the four warm years 1990, 1993, 1995, and 1997; note that wind
patterns (based on SLP) in both March and April impact April temperature anomalies. There was
anomalously low SLP over much of the Arctic with anomalous southwesterly winds in
opposition to the climatological late winter pattern of cold easterlies in the region from eastern
Siberia east to northeastern Canada. There are also southerly wind anomalies feeding warm air
north of Siberia. The previous warm spring period in Beringia from 1939–1941 (Fig. 11, left)
14
does not show lower SLP anomalies over the Arctic. The major spring warming in North
America appears to be associated with mid-latitude weather systems. The 1953–55 SLP (Fig. 11,
center) does show lower pressures in the central Arctic, again with warm anomalies in Siberia,
Beringia, and eastern Canada. However, in 1953–55 the main center of action in SLP shifted to
the western Arctic and away from the Atlantic, in contrast to the 1990s fields and the classical
AO definition (Thompson and Wallace 1998). Mid 20th century spring temperature anomaly
fields can be understood as resulting from subarctic advection fields and are not part of
hemispheric-wide Arctic changes similar to those of the 1990s (Shindell 2003).
In summary, there are five features shown in the visual inspection and PCA for spring:
the generally cool temperature anomalies before 1920 with some local variability, the regional
warming episodes near 1939–1941 and 1953–1955, the warm Siberian and Canadian
temperatures in the 1970s, and the unique warm pan-Arctic temperatures of the 1990s.
c. Summer, fall, and other months
For summer, one would expect smaller anomalies in part because of increased importance
of radiative processes and the melting of sea ice which buffers temperature extremes near coastal
stations. In northern Europe summer temperature anomalies are often reduced because the land-
sea contrast is less compared to winter. Hence note that the temperature anomaly scale on Fig. 12
for summer (July–August) is half that for the winter (Fig. 2) and spring (Fig. 8). Figure 12 shows
that there are generally cool periods before 1910 in northern Europe (Stns. 1–7) and before 1970
in much of Siberia eastward to northwestern Canada (Stns. 20–40). The strength of the warm
anomalies in the 1930s through the 1960s from Baffin Bay eastward through western Siberia
(Stns. 51–20) and cool anomalies for the rest of the Arctic are similar to those in winter data
(Fig. 2) but are unlike those in the preceding spring. However, summer temperature anomalies in
recent decades mirror spring warm anomalies from Siberia eastward to Canada (Stns. 16–49).
PCAs of summer temperature anomalies are not shown as they had little large-scale temporal or
15
spatial structure, presumably because of the small amplitudes and more local character compared
with winter or spring.
Like spring the temperature anomalies of fall, as represented by October (Fig. 13), show
strong longitudinal Arctic-wide covariability. Particularly strong are the warm anomalies from
the mid 1930s to the early 1950s. This pattern is arguably repeated in the 1980/1990s but with
more temporal variability at individual stations. East Greenland eastward through northern
Europe (Stns. 54–7) was generally cool before 1930.
We have combined two months with similar anomaly patterns for winter (December–
January) and summer (July–August) and have used one month to represent spring and fall
transitions. The following is a brief description of how the other six months relate to those
shown; monthly plots are available at www.unaami.noaa.gov. Our four seasonal figures 2, 8, 12,
13 are in general representative of the variations for all months. February is similar to
December/January except for a warming in north America in recent years. March is similar to
April, only the intensity of the pattern is weaker. May and June are similar to April with cold
historical temperatures in northern Europe and warm temperatures across the Arctic in recent
years. One difference is warmer historical temperatures in east Greenland in May. September and
November are much like October with warm Siberian and east Greenland temperature anomalies
from the late 1930s to early 1950s. In September there are warm anomalies in Baffin Bay and
East Greenland regions in the 1920s.
4. Discussion
a. Comparison with previous PCAs
All previous PCAs that we are aware of have been based on gridded fields of SAT. Walsh
(1977) conducted a PCA analysis north of 60°N for 1955–1975 from temperature anomalies for
all individual months. His first mode is the N. Atlantic seasaw and his second mode has a central
Siberia and northern N. America in phase, which echos our winter pattern C. Kelley et al. (1982)
uses annual data north of 60°N for 1881–1980. Their first mode is an in phase behavior between
16
Baffin Bay and Northern Siberia with the PC peaking in the late 1920s and 1930. Their second
mode is a N. Atlantic seasaw. Semenov and Bengtsson (2003) use six month averages NDJFMA
north of 40°N for 1892–1999. Their first mode is the N. Atlantic seasaw, and its PC becomes
positive after 1970. Their second mode is a positive-negative hemispheric pattern with a high
frequency character and potential ENSO influence. Their third mode shows a mid-century
1920–1955 warm period with Baffin Bay and northern Europe in phase.
All analyses include the winter seasaw pattern as an early detected mode. Our winter
pattern A time series stay interdecadal in character during the recent decade, while the Semenov
and Bengtsson (2003) first PC shows an upward trend during the 1980–1990s. Perhaps their
compositing of March–April, which does have this trend, with winter months contributes to this
result. Our second winter pattern B and first spring pattern A show mid-century warming events
in the 1920–1930s and 1940–1950s, respectively. The Semenov and Bengtsson (2003) mid-
century third PC also seems to composite these two seasonal events. The Kelly et al. (1982)
results of an in phase behavior between Baffin Bay and Siberia in the 1930s and our winter
pattern B support the conclusion of Semenov and Bengtsson (2003) that the character of the late
1930s is separate from the NAO, at least based on PCA mathematics.
b. Multidecadal processes 1800–2002
Except for winter in northern Europe and fall in central Siberia, all SAT time series show
a general warming over the instrumental record (Figs. 2, 8, 12, 13). For comparison to the 19th
century several references are of note in addition to the Stockholm record. Gervais and
MacDonald (2001) state that trees on the Kola peninsula show suppressed growth for a twenty-
year period after the 1809 eruption. Lee et al. (2000) reports cold springs in Finland before 1900.
It is likely that the Arctic was cold in certain decades in the 19th century with a warming rebound
in the first half of the 20th century. That these trends are often in non-winter months is consistent
with papers that suggest a reduction in volcanic influence from the mid 1800s through the 1950s.
However, the upward trend of SAT has continued during recent decades despite an increase in
17
volcanic forcing. Several authors suggest that this continued trend is due to anthropogenic
forcing (d’Arrigo and Jacoby 1993; Free and Robock 1999, Fig. 7; Stott et al. 2001).
While cooling due to increased volcanic aerosols may be true for the non-winter months,
several authors make a case for winter warming in Eurasia from an increase in volcanic eruptions
(Robock and Mao 1992; Graf et al. 1993; Stenchikov et al. 2002). Robock and Mao (1992) note
high latitude winter warming from the 12 largest volcanoes since 1883. The physical argument is
that the radiational effects are large at low latitude, producing primarily a dynamical response in
mid and high latitudes due to increased latitudinal temperature gradients. Thus the winter
warming of the European sector in the 1930s, which influences the pan-Arctic annual average
temperature anomalies, cannot be clearly attributed to the lack of vulcanism. The winter warming
in Europe in the 1930s–early 1940s appears to be more of a high latitude internal variability
event following a warm phase of the North Atlantic seasaw in the 1920s.
While there is a general minimum in temperature anomalies in the Arctic during the
1960–1970s, we see no clear evidence for a 50-year Low Frequency Oscillation (LFO) in SAT as
proposed by Polyakov et al. (2002), in our regional/seasonal analyses of the instrumental record,
the Stockholm wavelet analysis, or proxy data before 1920 (Briffa and Osborn 2002). That the
physics for the mid-century warming may be different from the 1990s warm period (Hanssen-
Bauer and Førland 1998) is an additional argument against an LFO. Thus, there is no clear
justification for extrapolating a 50-year cycle forward, as there is no clear extrapolation backward
to the 1800s.
Temperature anomalies in the Arctic, at least in fall through spring, are primarily driven
through temperature advection. This is documented for the Scandinavian sector in winter (Fu et
al. 1999) and the remaining Arctic in spring (Overland et al. 2002). The positive phase of the AO
in the winter provides warming and cooling in the Atlantic sector but not a strong temperature
advection signal in the remainder of the Arctic (Rigor et al. 2002). Considerable advective
warming does occur in the remainder of the Arctic in the spring when the polar vortex breaks
down. The strong dynamic (wind) control in the Arctic cool season suggests care should be taken
18
in estimating annual temperatures from summer-based proxy data which are often controlled by
radiational processes. In assessing causality for Arctic change we should not only look to local
radiative processes but also to lower latitude radiative processes which force the subarctic
circulation through changes in latitudinal temperature gradients (Stenchikov et al. 2002). It is
also possible that sea ice processes, change in land cover such as the increase in shrubs (Sturm et
al. 2001), and ozone chemistry (Newman et al. 2001) promote the persistence of spring and
summer circulation anomalies on decadal scales. Such a decadal feedback process could have
occurred in the late 1930s (Bengtsson et al. 2003).
5. Conclusions
There are considerable differences in decadal SAT variability in the Arctic across seasons
and regions. These differences are apparent in both the visual and semi-quantitative (PCA)
investigation of the data. Much previous work centers on annual, six month, and Arctic-wide
averages (Kelley et al. 1982; Przybylak 2000; Polyakov et al. 2002; Semenov and Bengtsson
2003) which lead to considerable confounding of Arctic processes.
In hindsight our separating the PCA into three periods 1886–2002, 1912–2002, and
1936–2002 to understand the influence of historical station coverage is justified. The second
EOF for the more recent 1936–2002 period was, for example, different from the other PCAs
based on longer data periods. Using the larger amounts of recent station data to establish an EOF
pattern, and then extrapolating to earlier time periods by projecting onto fewer stations to
determine the PCs, is questionable.
In this study we have used PCA to track the major features shown in time/longitude plots
of SAT. For example, the nearly Arctic-wide simultaneous events of the 1910s and 1980s are
resolved by our winter pattern C. Our PCs differ and support previous multi-month composite
analyses. Our winter pattern A emphasizes a continuing NAO interdecadal oscillation and a
separate longer duration warming event is represented by pattern B. Our spring pattern A
emphasizes warming in the 1940s/1950s and 1990s. However, based on the preponderance of the
19
evidence only spring in the 1990s and possibly the summer show an Arctic-wide SAT signal,
consistent with a recent pan-Arctic change in circulation patterns. It will be important to monitor
this large-scale change over the next several decades.
Acknowledgments. We appreciate discussions with K. Wood and N. Bond on aspects of this
paper. We thank NSF through the SEARCH Project Office for support of this study. The paper
is also a contribution to SEARCH through the NOAA Arctic Research Office. JISAO
contribution number 997 under NOAA Cooperative Agreement NA17RJ1232.
20
REFERENCES
Ahlmann, H. W., 1948: The present climate fluctuation. Geogr. J., 112, 165–195.
Ambaum, M. H. P., and B. J. Hoskins, 2002: The NAO troposphere-stratosphere connection. J.
Climate, 15, 1969–1978.
Bengtsson, L., V. A. Semenov, and O. Johannessen, 2003: The early century warming in the
Arctic—a possible mechanism. Report 345 Max Planck Institut fur Meteorologie, ISSN
0937-1060, 31 pp.
Briffa, K. R., and T. J. Osborn, 2002: Blowing hot and cold. Science, 295, 2227–2228.
Chapman, W. L., and J. E. Walsh, 1993: Recent variations of sea ice and air temperature in high
latitudes. Bull. Am. Meteorol. Soc., 74, 33–47.
Crowley, T. J., 2000: Causes of climate change over the past 1000 years. Science, 289, 270–277.
d’Arrigo, R. D., and G. C. Jacoby, 1993: Secular trends in high northern latitude temperature
reconstructions based on tree rings. Clim. Change, 25, 163–177.
Davis, R. E., 1976: Predictability of sea temperature and sea level pressure anomalies over the
North Pacific Ocean. J. Phys. Oceanogr., 6, 249–266.
Free, M., and A. Robock, 1999: Global warming in the context of the little ice age. J. Geophys.
Res., 104, 19 057–19 070.
Fu, C., H. Diaz, D. Dong, and J. O. Fletcher, 1999: Changes in atmospheric circulation over
northern hemisphere oceans associated with the rapid warming of the 1920s. Int. J.
Climatol., 19, 581–606.
Gervais, B. R., and G. M. MacDonald, 2001: Tree-ring and summer-temperature response to
volcanic aerosol forcing at the northern tree-line, Kola Peninsula, Russia. The Holocene,
11, 499–505.
Graf, H. F., I. Kirchner, A. Robock, and I. Schult, 1993: Pinatubo eruption winter climate effects:
model versus observation. Clim. Dyn., 9, 81–93.
21
Hanssen-Bauer, I., and E. J. Førland, 1998: Long-term trends in precipitation and temperature in
the Norwegian Arctic: can they be explained by changes in atmospheric circulation
patterns? Clim. Res., 10, 143–153.
IPCC Report, Houghton, J. T. and Coauthors, 2001: Climate Change 2001: The scientific basis.
Cambridge Press, 881 pp.
Johannessen, O. M., and Coauthors, 2003: Arctic climate change—observed and modeled
temperature and sea ice. Nansen Center Technical Report, 26 pp.
Kelly, D. M., and Coauthors, 1982: Variations in surface air temperatures: Part 2, Arctic regions
1881–1980. Mon. Weather Rev., 110, 71–83.
Lee, S. E., M. C. Press, and J. A. Lee, 2000: Observed climate variations during the last 100
years in Lapland, Northern Finland. Int. J. Climatol., 20, 329–346.
Moberg, A., H. Bergström, J. R. Krigsman, and O. Svanered, 2002: Daily air temperature and
pressure series for Stockholm (1756–1998). Clim. Change, 53, 171–212.
Moritz, R. E., C. M. Bitz, E. J. Steig, 2002: Dynamics of recent climate change in the Arctic.
Science, 297, 1497–1502.
New, M., M. Hulme, and P. Jones, 1999: Representing twentieth-century space-time climate
variability. Part I: Development of a 1961–90 mean monthly terrestrial climatology. J.
Climate, 12, 829–856.
New, M., M. Hulme, and P. Jones, 2000: Representing twentieth-century space-time climate
variability. Part II: Development of 1901–96 monthly grids of terrestrial surface
temperature. J. Climate, 13, 2217–2238.
Newman, P. A., E. R. Nash, and J. E. Rosenfield, 2001: What controls the temperature of the
Arctic stratosphere in spring? J. Geophys. Res., 106, 19 999–20 010.
Omstedt, A., and D. Chen, 2001: Influence of atmospheric circulation on the maximum ice
extent in the Baltic Sea. J. Geophys. Res., 106, 4493–4500.
Overland, J. E., M. Wang, and N. A. Bond, 2002: Recent temperature changes in the western
Arctic during spring. J. Climate, 15, 1702–1716.
22
—, M. C. Spillane, and N. N. Soreide, 2004: Integrated analysis of physical and biological pan-
Arctic change. Clim. Change, April, in press.
Percival, D. B., and H. O. Mofjeld, 1997: Analysis of subtidal coastal sea level fluctuations using
wavelets. J. Amer. Stat. Assoc., 92, 868–880.
Polyakov, I. V., and Coauthors, 2002: Observationally based assessment of polar amplification
of global warming. Geophys. Res. Lett., 29, doi:10.1029/2001GL011111.
Przybylak, R., 2000: Temporal and spatial variation of surface air temperature over the
instrumental observations in the Arctic. Int. J. Climatol., 20, 587–614.
—, 2002: Variability of Air Temperature and Atmospheric Precipitation in the Arctic. Kluwer,
330 pp.
—, 2003: The Climate of the Arctic. Kluwer, 270 pp.
Rigor, I. G., R. L. Colony, and S. Martin, 2000: Variations in surface air temperature
observations in the Arctic 1979–1997. J. Climate, 13, 896–914.
—, J. M. Wallace, and R. L. Colony, 2002: Response of sea ice to the Arctic Oscillation. J.
Climate, 15, 2648–2663.
Robock, A., 2000: Volcanic eruptions and climate. Rev. Geophys., 38, 191–219.
—, and J. Mao, 1992: The volcanic signal in surface temperature observations. J. Climate, 8,
1086–1103.
Rogers, J. C., 1985: Atmospheric circulation changes associated with the warming over the
northern North Atlantic in the 1920s. J. Climate Appl. Meteorol., 24, 1303–1310.
—, and E. Mosley-Thompson, 1995: Atlantic cyclones and the mild Siberian winters of the
1980s. Geophys. Res. Lett., 22, 799–802.
—, and M. J. McHugh, 2002: On the separability of the North Atlantic Oscillation and the Arctic
Oscillation. Clim. Dyn., 19, 599–608.
Semenov, V. A., and L. Bengtsson, 2003: Modes of the wintertime Arctic temperature
variability. Geophys. Res. Lett., 30, 10.1029/2003GL017171.
23
Serreze, M. C., and Coauthors, 2000: Observational evidence of recent change in the northern
high-latitude environment. Clim. Change, 46, 159–207.
Skeie, P., 2000: Meridional flow variability over the Nordic seas in the Arctic Oscillation
framework. Geophys. Res. Lett., 27, 2569–2572.
Shindell, D., 2003: Whither Arctic climate. Science, 299, 215–216.
Stenchikov, G., and Coauthors, 2002: Arctic Oscillation response to the 1991 Mount Pinatubo
eruption: effects of the volcanic aerosols and ozone depletion. J. Geophys. Res., 107,
doi:10.1029/2002JD002090.
Stott, P. A., and Coauthors, 2001: Attribution of twentieth century temperature change to natural
and anthropogenic causes. Clim. Dyn., 17, 1–21.
Sturm, M., C. Racine, and K. Tape, 2001: Increasing shrub abundance in the Arctic. Nature, 411,
546–547.
Thompson, D. W. J., and J. M. Wallace, 1998: The Arctic Oscillation signature in the wintertime
geopotential height and temperature fields. Geophys. Res. Lett., 25, 1297–1300.
Trenberth, K. E., and D. A. Paolino, Jr., 1980: The Northern Hemisphere sea-level pressure data
set: Trends, errors and discontinuities. Mon. Wea. Rev., 108, 855–872.
van Loon, H., and J. C. Rogers, 1978: The seasaw in winter temperatures between Greenland and
northern Europe, Part 1: General description. Mon. Weather Rev., 106, 296–310.
Walsh, J. E., 1977: The incorporation of ice station data into a study of recent Arctic temperature
fluctuations. Mon. Weather Rev., 105, 1527–1535.
24
Figure Captions
1. Location map for 59 stations used in the study (See Table 1). Also indicated are approximate
regional sectors: northern Europe, Siberia, Beringia, northwestern Canada, Baffin Bay, and E
Greenland. Smaller station numbers in italics denote stations used in the time/longitude plots but
excluded from the PCA analysis due to short records or geographic proximity.
2. Time/longitude plot of five-year averaged surface air temperature (SAT) anomalies for winter
(December–January) relative to the 1961–1990 mean for each station. Note the evidence of an
NAO interdecadal seesaw response in northern Europe, warm Siberian temperatures in the
1940s, and more hemispheric warming in the 1980s. Temperatures were generally cold before
1920 outside of northern Europe.
3. Winter pattern A is represented by the PCA analysis with records beginning in 1886, 1912,
and 1936 (Table 2). The Empirical Orthogonal Functions (EOFs) are shown as amplitudes at
station locations as a function of longitude. The three EOFs show a similar seesaw pattern with
northern Europe out of phase with the Baffin Bay region. The lower panel shows the Principal
Component (PC) time series. Different lines (solid, dashed) represent the three starting times for
the different realizations. The PCs do not have a long-term trend.
4. Polar stereographic plots of SAT and sea level pressure anomalies for December–January
1933, 1936, and 1983–1984. The years correspond to the month of January. SAT data are from
Climate Research Unit CRU T2.0 and the SLP data are from the Trenberth set at NCAR.
5. Winter pattern B based on EOF mode 2 for data beginning in 1886 and EOF mode 3 for data
beginning in 1912. All patterns show an in phase behavior between Baffin Bay and the northern
Europe.
25
6. Winter pattern C is represented by EOF mode 2 for station data beginning 1912 and EOF
mode 3 for data starting 1936. Note the broad positive signal with the exception of Baffin Bay.
7. Multi-resolution analysis of the December–January monthly temperature record for
Stockholm (bottom curve) based on the Haar maximal overlap discrete wavelet transform with
symmetric boundary conditions; the time series is from Moberg et al. (2002). Note the weak
upward trend in the low frequency curve (top) and the considerable energy in the interdecadal (8-
year half cycle) decomposition during the 20th century.
8. Time/longitude plot of temperature anomalies for spring (April) similar to Fig. 2. Note the
strong spatial covariability of the fields with warm hemispheric temperatures in the 1990s and
cool temperatures before 1920.
9. Spring April pattern A is fit by the EOF mode 1 for the series beginning in 1886, 1912, and
1936. There is a general in-phase Arctic-wide behavior with the weakest impact in the North
Atlantic. The PC shows the warm anomalies in the 1950s and 1990s.
10. Spring pattern B supported by EOF mode 2 for the three analysis periods, shows an out of
phase behavior between northern Europe and the Beringia/North America regions.
11. Comparison maps of the April SAT anomalies and March–April SLP fields for 1939–41,
1953–55, and the warm years of the 1990s. While the temperature anomaly patterns are
somewhat similar, their causes relate to different advective fields. The temperature fields are
from the CRU T2.0 and the SLP fields from NCAR.
12. Time/longitude plots of temperature anomalies for summer (July–August) similar to Fig. 2.
Note that the range of the color scale is half of the winter and spring plots (Figs. 2, 8).
26
13. Time/longitude plots of temperature anomalies for fall (October) similar to Fig. 2.
27
Table 1. Surface Air Temperature (SAT) stations, north of 64°N, at which lengthy monthly data
records are available. The percent data and time span statistics are based on the entire record at
the station and may vary between months.
Zonal WMO East North Time Span Percent
Order Station ID Lon. Lat. Data Locale
1 Bodo Vi 01152 14.4 67.3 1868:2002 100.0 Water
2 Svalbard* 01008 15.5 78.3 1912:2002 97.8 Polar desert
3 Tromso 01025 18.9 69.7 1856:2002 100.0 Tundra
4 Bjornoya 01028 19.0 74.5 1949:2002 99.5 Water
5 Karesuando 02080 21.5 68.5 1881:2002 100.0 Wooded tundra
6 Haparanda 02196 24.1 65.8 1860:2002 99.9 Coastal
7 Vardo 01098 31.1 70.4 1829:2002 87.9 Water
8 Ostrov Victoria 20026 36.8 80.2 1959:1995 93.2 Water
9 Archangelsk 22550 40.7 64.5 1813:2002 80.4 Main taiga
10 Kanin Nos 22165 43.3 68.7 1915:2002 94.3 Water
11 Nagurskoye 20034 47.6 80.8 1952:1995 98.9 Water
12 Malye Karmakuly 20744 52.7 72.4 1897:1999 90.6 Tundra
13 Narjan-Mar 23205 53.0 67.6 1926:2002 97.7 Tundra
14 Ostrov Rudolfa 20049 58.0 81.8 1932:1995 85.8 Water
15 Mys Zelanija 20353 63.6 77.0 1931:1996 98.7 Polar desert
16 Salehard 23330 66.7 66.5 1882:2002 98.6 Bogs/woods
17 Ostrov Belyj 20667 70.1 73.3 1933:2001 97.1 Water
18 Ostrov Dikson 20674 80.4 73.5 1916:2002 96.6 Water
19 Turuhansk 23472 87.9 65.8 1880:2002 95.7 Northern taiga
20 Mys Golomianny 20087 90.6 79.6 1930:2002 90.6 Water
21 Tura 24507 100.2 64.3 1928:2002 97.9 Main taiga
22 Khatanga 20891 102.5 72.0 1929:2002 92.6 Wooded tundra
23 O. Preobrazhenia 21504 112.9 74.7 1934:1996 98.3 Water
24 Bukhta Tiksi 21824 128.9 71.6 1932:2002 96.9 Water
25 Verhojansk 24266 133.4 67.6 1885:2002 96.5 Northern taiga
26 Ostrov Kotelnyj 21432 137.9 76.0 1933:2002 91.7 Polar desert
27 Mys Salaurova 21647 143.2 73.2 1928:2001 96.1 Water
28 Zyrjanka 25400 150.9 65.7 1935:2002 98.8 Northern taiga
29 Bukhta Ambarchik 25034 162.3 69.6 1935:1999 98.8 Water
30 Anadyr 25563 177.6 64.8 1898:2002 92.0 Tundra
31 Mys Shmidta 25173 180.6 68.9 1932:2002 95.0 Tundra
32 Ostrov Vrangel 21982 181.5 71.0 1926:2002 96.9 Water
33 Bukhta Providenia 25594 186.8 64.4 1934:2002 96.0 Tundra
34 Mys Uelen 25399 190.2 66.2 1918:2002 84.6 Water
35 Nome 70200 194.6 64.5 1906:2002 98.7 Water
36 Barrow 70026 203.2 71.3 1893:2002 100.0 Tundra
37 Fairbanks 70261 212.1 64.8 1904:2002 98.7 Main taiga
38 Inuvik 71957 226.5 68.3 1892:2002 84.5 Tundra
28
Table 1. (continued).
Zonal Station WMO East North Time Span Percent Locale
Order ID Lon. Lat. Data
39 Norman Wells 71043 233.2 65.3 1898:2002 95.2 Main taiga
40 Sachs Harbour 71051 234.7 72.0 1955:2002 94.4 Water
41 Mould Bay 71072 240.7 76.2 1948:2000 99.1 Water
42 Coppermine 71938 244.9 67.8 1930:2002 96.6 Tundra
43 Cambridge Bay 71925 254.9 69.1 1929:2002 87.7 Tundra
44 Baker Lake 71926 263.9 64.3 1946:2002 94.0 Tundra
45 Resolute 71924 265.0 74.7 1947:2002 97.9 Tundra
46 Eureka 71917 274.1 80.0 1947:2002 97.9 Water
47 Pond Inlet 71095 282.0 72.7 1922:2002 77.4 Tundra
48 Cape Dorset 71910 283.5 64.2 1927:2002 83.7 Water
49 Frobisher Bay 71909 291.4 63.8 1913:2002 89.0 Tundra
50 Alert 71082 297.7 82.5 1950:2002 96.4 Polar desert
51 Upernavik 04210 303.8 72.8 1873:2002 98.8 Water
52 Godthab Nuuk 04250 308.2 64.2 1866:2002 98.0 Tundra
53 Egedesminde† 04220 308.9 69.2 1866:2002 98.2 Water
54 Angmagssalik 04360 322.4 65.6 1895:2002 98.5 Water
55 Kap Tobin 04339 338.0 70.5 1931:2002 64.8 Water
56 Stykkisholmur 04013 337.3 65.1 1846:2002 96.6 Tundra
57 Reykjavik 04030 338.1 64.1 1901:2002 99.7 Grass
58 Station Nord 04312 343.3 81.6 1952:2002 77.5 Ice
59 Jan Mayen 01001 351.3 70.9 1921:2002 99.0 Water
*The Svalbard series is an amalgam of Istfjord Radio (WMO#01005 1912–1980) and the
continuing Svalbard Lufthavn record, 1977–Present.
†The Egedesminde series combines Jakobshavn (WMO#04221 1866–1980) data with the
continuing Egedesminde record, 1949–Present.
29
Table 2. Winter (Dec–Jan) patterns and contributing EOF modes for the three separate analyses
with different record lengths.
Record Period
Description 1880–2002 1912–2002 1936–2002
Pattern A N Atlantic seesaw mode 1 mode 1 mode 1
Pattern B Baffin Bay/Siberia in phase mode 2 mode 3
Pattern C 1980s warming mode 2 mode 3
Figure 1
Figure 2
Figure 3
Figure 4
Figure 5
-7
-1
5
1750 1800 1850 1900 1950 2000
year
-2
0
2
-4
0
4
-2
0
2
-2
0
2
-2
0
2
-1
0
1
 
 
 
 
 

S
5
( 64 years)
D
5
(32 - 64 years)
D
4
(16 - 32 years)
D
3
(8 -16 years)
D
2
(4 - 8 years)
D
1
(2 - 4 years)
Nominal Pass
Band
Figure 7
Figure 8
Figure 9
Figure 10
1939-41
1953-55
1990,93,95,97
4.0
3.0
2.0
1.0
-1.0
-1.0
-2.0
-3.0
-1.0
-2.0
1.0
2.0
1.0
Composite of April SAT Anomalies (
o
C)
Composite of Mar-Apr SLP Anomalies (hPa)
Figure 11
Figure 12
Figure 13