ArticlePDF Available

Arctic Environmental Change of the Last Four Centuries

Authors:

Abstract and Figures

A compilation of paleoclimate records from lake sediments, trees, glaciers, and marine sediments provides a view of circum-Arctic environmental variability over the last 400 years. From 1840 to the mid-20th century, the Arctic warmed to the highest temperatures in four centuries. This warming ended the Little Ice Age in the Arctic and has caused retreats of glaciers, melting of permafrost and sea ice, and alteration of terrestrial and lake ecosystems. Although warming, particularly after 1920, was likely caused by increases in atmospheric trace gases, the initiation of the warming in the mid-19th century suggests that increased solar irradiance, decreased volcanic activity, and feedbacks internal to the climate system played roles.
Content may be subject to copyright.
Arctic Environmental Change of the
Last Four Centuries
J. Overpeck,* K. Hughen, D. Hardy, R. Bradley, R. Case, M. Douglas, B. Finney, K. Gajewski,
G. Jacoby, A. Jennings, S. Lamoureux, A. Lasca, G. MacDonald, J. Moore,
M. Retelle, S. Smith, A. Wolfe, G. Zielinski
A compilation of paleoclimate records from lake sediments, trees, glaciers, and marine
sediments provides a view of circum-Arctic environmental variability over the last 400
years. From 1840 to the mid-20th century, the Arctic warmed to the highest temperatures
in four centuries. This warming ended the Little Ice Age in the Arctic and has caused
retreats of glaciers, melting of permafrost and sea ice, and alteration of terrestrial and
lake ecosystems. Although warming, particularly after 1920, was likely caused by in-
creases in atmospheric trace gases, the initiation of the warming in the mid-19th century
suggests that increased solar irradiance, decreased volcanic activity, and feedbacks
internal to the climate system played roles.
Global climate change is likely amplified
in the Arctic by several positive feedbacks,
including ice and snow melting that de-
creases surface albedo, atmospheric stability
that traps temperature anomalies near the
surface, and cloud dynamics that magnify
change (1–3). The Arctic, in turn, influ-
ences climate change at lower latitudes
through changes in river runoff and effects
on global thermohaline circulation, impacts
on atmospheric circulation, and modulation
of atmospheric CO
2
and CH
4
concentra-
tions (1, 4). Although the Arctic is now
one of the least disturbed regions on Earth,
it may also be one of the most susceptible to
both natural and human-induced climate
change.
The instrumental record of Arctic cli-
mate change is brief and geographically
sparse. The few records that extend back to
the beginning of the 20th century suggest
that the Arctic has warmed by about 0.6°C
since that time, with peak temperatures
(1.2°C more than in 1910) around 1945
(2). Although some areas of the Arctic
have cooled recently—for example, as a
result of dynamics within the North Atlan-
tic (5)—the observed 20th-century Arctic-
average temperature increase exceeds that
of the hemisphere as a whole. The record is
thus consistent with amplification of cli-
mate change through snow/ice radiation
and other feedbacks (6, 7).
In this article, we use the paleoenviron-
mental record to assess the climate events
of this century from the perspective of the
last four centuries. We build on previous
work (8–10) by compiling a variety of com-
plementary paleoenvironmental indicators
of climate from around the entire Arctic.
This perspective permits the visualization of
natural subdecadal to century-scale climate
variability in the circum-Arctic region and
allows us to examine the role that natural
forcing mechanisms play in driving Arctic
climate. Given the growing focus on Arctic
environmental dynamics (4), we also exam-
ine the centuries-long paleoenvironmental
perspective to determine whether the cli-
mate changes we infer have resulted in
changes to the natural Arctic ecosystem.
The Paleoclimate Perspective
Proxy data from lakes, wetlands, ice cores,
and marine sources demonstrate that Arctic
J. Overpeck is at the National Oceanographic and Atmo-
spheric Administration (NOAA)–National Geophysical
Data Center (NGDC) Paleoclimatology Program, 325
Broadway, Boulder, CO 80303, USA, and Institute of
Arctic and Alpine Research (INSTAAR), University of Col-
orado, Boulder, CO 80309, USA. K. Hughen, A. Jen-
nings, J. Moore, and A. Wolfe are at INSTAAR, University
of Colorado, Boulder, CO 80309, USA. D. Hardy, R.
Bradley, and S. Smith are in the Department of Geo-
sciences, University of Massachusetts, Amherst, MA
01003, USA. R. Case and G. MacDonald are in the De-
partment of Geography, University of California, Los An-
geles, CA 90095, USA. M. Douglas is in the Department
of Geology, University of Toronto, Toronto, Ontario M5S
3B1, Canada. B. Finney is at the Institute of Marine Sci-
ence, University of Alaska, Fairbanks, AK 99775, USA. K.
Gajewski is in the Department of Geography, University of
Ottawa, Ottawa, Ontario K1N 6N5, Canada. G. Jacoby is
at the Lamont-Doherty Earth Observatory, Columbia Uni-
versity, Palisades, NY 10964, USA. S. Lamoureux is in
the Department of Earth and Atmospheric Sciences, Uni-
versity of Alberta, Edmonton, Alberta T6G 2H3, Canada.
A. Lasca and M. Retelle are in the Geology Department,
Bates College, Lewiston, ME 04204, USA. G. Zielinski is
at the Climate Change Research Center, University of
New Hampshire, Durham, NH 03824, USA.
*To whom correspondence should be addressed. E-mail:
jto@ngdc.noaa.gov
Table 1. Paleoclimate site information.
Map no. Site name Proxy type Source
1 Svalbard Ice Core Percent summer melt (69)
2 Tornetrask Tree ring widths (22, 28)
3 Polar Urals Tree ring widths (22, 70)
4 Lena River Tree ring widths This article
5 Kolyma (Jack London Lake) Tree ring widths (71)
6 Site 412 Tree ring widths (8, 9)
7 Arrigetch Peaks Tree ring widths (8, 9)
8 Dune Lake
13
C isotopes This article
9 Sheenjek Tree ring widths (8, 9)
10 TTHH Tree ring widths (8, 9)
11 Eagle Fecal Tree ring widths (8, 9)
12 Franklin Mountains Tree ring widths (8, 9)
13 MacKenzie Mountains Tree ring widths (8, 9)
14 Coppermine Tree ring widths (8, 9)
15 Hornby Cabin Tree ring widths (8, 9)
16 Churchill Tree ring widths (8, 9)
17 Lake DV09 Varve thicknesses (63)
18 Devon Island Ice Core Percent summer melt (27)
19 Lake C2 Varve thicknesses (21)
20 Lake C3 Varve thicknesses (65)
21 Lake Tuborg Varve thicknesses This article
22 Castle Peninsula Tree ring widths (8, 9)
23 Kuujuaq Tree ring widths (8)
24 Upper Soper Lake Dark lamination thicknesses (20)
25 Okak Tree ring widths This article
26 Salt Water Pond Tree ring widths (72)
27 Donard Lake Varve thicknesses This article
28 South Greenland Ice Core Percent summer melt (10, 73)
29 Nansen Fjord Fossil foraminifera (29)
ARTICLE
www.sciencemag.org zSCIENCE zVOL. 278 z14 NOVEMBER 1997 1251
interannual to century-scale environmental
variability of the last 400 years is superim-
posed on longer-term changes of the Holo-
cene (the last 10,000 years). Most of the
Arctic experienced summers warmer (1° to
2°C) than today during the early to middle
Holocene, but the timing of this Milanko-
vitch-driven warmth (that is, warmth re-
sulting from greater summer insolation) dif-
fered geographically because of the effects
of local sea surface temperatures (SSTs) and
land-ice cover (11–13). Decreases in sum-
mer insolation—perhaps coupled with oth-
er, more abrupt changes in climate forc-
ing—led to successively cooler summers in
the late Holocene; this trend culminated in
the Little Ice Age, a period that began
before 1600 (14–17) and evidently ended
sometime in the last century (11, 12, 14).
Our compilation of paleoclimate records
provides a 400-year view of Arctic variabil-
ity that is superimposed on lower frequency
Holocene trends. This view is based (18)on
more than 20 records from the North
American Arctic and substantially fewer
records spread out across the Eurasian Arc-
tic (Table 1 and Figs. 1 and 2). Although
covariation among continental records in
Alaska and western Canada (Fig. 2A) sug-
gests that the Eurasian records may be
broadly representative, a need for caution is
highlighted by the differences among
records from the eastern Canadian Arctic
and Greenland (Fig. 2B), a region marked
by geographic variability even in the instru-
mental period (2). Thus, although our com-
pilation provides a much improved basis for
understanding the patterns and causes of
interannual to century-scale environmental
variability in the circum-Arctic, future
work will be needed to increase data cover-
age, particularly in Eurasia.
The most notable pattern of change re-
vealed by most records across the Arctic is
the near ubiquitous transition from anoma-
lously cold conditions of the 19th century
to peak warm conditions of the 20th cen-
tury. This event can be seen even more
clearly in a new proxy record of average
Arctic temperature change over the last
400 years (19) (Fig. 3). Where quantitative
estimates are available, it appears that the
19th-to 20th-century warming was to
3°C locally (8, 15, 20–25) and averaged
about 1.5°C across the Arctic. We thus
confirm that the 20th-century warming ob-
served in the instrumental record was only
part of a more circum-Arctic climate event
that began in the mid-19th century and
marked the end of the Little Ice Age in the
Arctic (8, 26, 27).
Another striking aspect of Arctic tem-
perature change over the last 400 years is
that most of the Arctic experienced cooling
in the first part of the 19th century, result-
ing in the coldest temperatures of the Little
Ice Age. In addition, although much of the
Arctic was colder than today during parts of
the 17th century, several of the records
show temperatures nearly as warm as today
during the 18th century (9). Arctic climate
change before 1800 was more regional in
nature than after this time. For example,
many sites around the Arctic were warmer
at some time between 1700 and 1820 than
they were later in the 19th century, but the
timing and duration of this warmth varied
from region to region, as did the preceding
17th-century period of generally colder
conditions.
The annually dated record of Arctic cli-
mate variability encompassing the last 1000
years has less spatial coverage than does the
multiproxy record of the last 400 years.
Sediment, ice core, historical, and tree ring
data for this earlier period indicate that
although Arctic summers of the 20th cen-
tury were generally the warmest of the last
400 years, they may not have been the
warmest of the last millennium (21, 28–
30). The few time series of climate change
spanning the last millennium also suggest
that the Arctic was not anomalously warm
throughout the so-called Medieval Warm
Period of the 9th to 14th centuries (31).
Climate Forcing Mechanisms
The paleoclimate record of the Arctic re-
veals large variability over all of the last 400
years. Natural variability in most regions
was as large before the 20th-century buildup
of atmospheric trace gases as afterward
(Figs. 2 and 3). Natural climate forcing
mechanisms must therefore be considered
before attributing some or all of the recent
changes in the Arctic to human influences.
The similarities of our new Arctic climate
reconstruction (Fig. 3) to earlier re-
constructions for the Northern Hemisphere
(8, 10), and the coincidence of high inferred
solar irradiance and warm Arctic tempera-
tures in both the 18th and 20th centuries,
suggest that solar forcing played a role (32–
36). The strong recovery of high tempera-
tures after the 1840s may also have been a
partial response to high solar irradiance at
that time. On the other hand, the lack of a
distinct prolonged cold period associated
with the 17th-to early 18th-century Maun-
der sunspot minimum period (34) argues
against a dominant role for solar forcing over
the past 400 years, as does the observation
that changes in solar irradiance evidently did
not lead climate warming into the 20th
century.
Comparison of our temperature record
with an Arctic volcanic sulfate record re-
cently developed from the Greenland Ice
Lake sediment records
Marine sediment records
Ice core records
Tree ring records
Fig. 1. Map showing locations of proxy climate records used in this study and listed in Table 1. Circles
around groups of sites indicate the groups averaged before calculating the Arctic-wide temperature
average. Isolated single sites not circled were averaged directly into the Arctic-wide series (19).
SCIENCE zVOL. 278 z14 NOVEMBER 1997 zwww.sciencemag.org1252
Sheet Project Two (GISP2) ice core (37,
38) indicates that most peaks in recon-
structed atmospheric volcanic sulfate load-
ing correspond to mean circum-Arctic cool-
ing (Fig. 3). The repeated coincidence of
high sulfate loading with the onset of Arc-
tic cold events suggests that eruptions en-
train positive ocean feedbacks capable of
enhancing and prolonging Arctic cooling.
For example, the anomalous early 19th-
century period of frequent large sulfur-pro-
ducing eruptions seems to have helped pre-
cipitate the drop into the coolest period of
the Arctic Little Ice Age (37, 39). These
findings agree with earlier evidence for vol-
canic forcing of Arctic temperatures over
the last 100 years (40), as well as the more
regional 200-year record of volcano-climate
linkages for sub-Arctic North America (41)
and earlier assertions that volcanic activity
plays a role in modulating the climate of the
Northern Hemisphere in general (10, 42).
It is difficult to attribute the lack of a clear
solar-climate correspondence, particularly
during the period of the Maunder sunspot
minimum (1650 to 1710), to counterbal-
ancing by volcanic forcing. However, it is
quite likely that the period of uncommonly
low volcanic activity from 1935 to 1960
may have contributed to the peak Arctic
summer temperatures at this time (43, 44),
reducing the implied potential sensitivity of
the climate system to solar forcing during
this same period (36).
Variability internal to Earth’s climate
system, particularly in the ocean’s thermo-
haline circulation, also can modulate cli-
mate over decades to centuries (45, 46).
The variability evident in our 400-year
Arctic compilation, although an order of
magnitude smaller in scale than that driven
by internal climate system processes over
the last deglaciation (47), may relate to
decadal-scale fluctuations in the state of the
North Atlantic Oscillation (NAO) or to
the transport of heat northward by thermo-
haline circulation. Such a forcing mecha-
nism would leave a diagnostic pattern most
obvious in, and downwind of, the North
Atlantic (46, 48). The Nansen Fjord record
from southeast Greenland (Fig. 2B), a proxy
for North Atlantic SSTs (49), indicates
that the range of North Atlantic variability
over the past 400 years exceeds that of the
recent instrumental record, and that the
last century was marked by warming in the
North Atlantic. However, comparison of
this long proxy record of North Atlantic
SSTs with terrestrial temperature proxies
from sites across the Arctic (Fig. 2) does not
reveal clear evidence for strong Atlantic
forcing of Arctic climate. The full role of
NAO and thermohaline forcing of Arctic
climate remains to be evaluated as more
multicentury time series from the circum–
North Atlantic region become available.
Half of the post-1840 warming (about
0.75°C) took place from 1840 to 1920,
during a period in which concentrations of
atmospheric CO
2
and CH
4
increased only
by about 20 ppm and 200 parts per billion
by volume (ppbv), respectively (Fig. 3).
Given the current best estimates of climate
sensitivity to trace-gas forcing [1.5° to 4.5°
global mean warming for a 280-ppm in-
crease in CO
2
, and less for a doubling of
CH
4
(6, 46)], trace-gas forcing alone might
be able to explain only a small part (0.1° to
0.4°C) of the pre-1920 warming. The first
half of the unprecedented 19th-to 20th-
century increase in Arctic temperatures ap-
pears to be a natural readjustment as volca-
nic forcing weakened, and irradiance (and,
to a lesser extent, greenhouse gases) in-
creased, between 1840 and 1920. After
1920, both anomalously high solar irradi-
ance and low volcanic aerosol loading likely
AB
1600 1700 1800 1900 2000 1600 1700 1800 1900 2000
Year
Europe Russia Alaska Western Canada
Eastern CanadaCentral Canada Greenland
Fig. 2. (A) Standardized 400-year
proxy climate records reflecting
surface air temperature for sites
from Arctic Europe east to western
Canada (Fig. 1 and Table 1) (18).
Red indicates temperatures greater
than one standard deviation warm-
er than average for the reference
period (1901–1960), whereas dark
blue indicates at least one standard
deviation colder than this average.
(B) Same as (A) but for sites in Can-
ada east to Greenland. All series are
presented as 5-year averages ex-
cept for sites 8 and 29, which are
plotted at their original lower reso-
lution. All time series represent sur-
face air temperature except for site
29, which represents SST. All time
series shown will be available at
http://www.ngdc.noaa.gov/paleo/
paleo.html.
ARTICLE
www.sciencemag.org zSCIENCE zVOL. 278 z14 NOVEMBER 1997 1253
continued to influence Arctic climate, but
exponentially increasing atmospheric trace-
gas concentrations probably played an in-
creasingly dominant role (36). More recent-
ly, the observed slowdown in warming from
1950 to 1970 (2) (Fig. 3) may have been
influenced by the increase in Arctic tropo-
spheric aerosols that occurred after 1950 (6,
50).
The Response to
Arctic Climate Variability
With the emergence of a coherent picture
of Arctic climate change over the last sev-
eral centuries comes the realization that
interannual to century-scale Arctic climate
variability is the norm. This variability has
affected many aspects of the Arctic envi-
ronment. The primary implication is that
today’s Arctic cryosphere (glaciers and fro-
zen ground) and biosphere (terrestrial,
lacustrine, and marine) are not at steady
state; they have changed and will continue
to change in response to an evolving Arctic
climate.
An obvious impact of recent climate
change is the widespread retreat of glaciers
throughout the Arctic over the last century
(51). On Baffin Island, retreats in glacial
equilibrium-line altitudes (52) coincided
with the 19th-to 20th-century warming
observed in sediment, tree ring, and ice core
records. Evidence for similarly large post–
Little Ice Age glacial retreats can also be
found farther east in Greenland, Iceland,
Spitsbergen, and Scandinavia (14, 53), as
well as to the north on Devon Island (27)
and to the west in Alaska (25). Although it
has been suggested that climate warming
may be accompanied by an increase in
snowfall sufficient to expand ice sheets
(54), it appears probable that most Arctic
glaciers will continue to melt if the Arctic
continues to warm, as forecast by model
simulations of the climate response to in-
creasing greenhouse gas concentrations (6,
55). This inference may not apply to the
Greenland Ice Sheet, whose modern mass
balance is poorly known (51).
High-latitude permafrost conditions have
also changed during the last 200 years.
Where available (that is, North America),
ground temperature records reconstructed
from borehole temperature logs support the
notion that large-scale warming has oc-
curred since the 19th century, but these
records also indicate that high-latitude per-
mafrost conditions have changed, and are
likely to continue changing, as the Arctic
warms (24). This future change in frozen
ground will, in turn, likely affect construc-
tion, transportation, hydrology, ecology,
and trace-gas fluxes in the Arctic (4).
The paleoenvironmental record of the
last four centuries supports recent ecologi-
cal studies (56) and indicates that, even if
climate warming takes place at the low end
of the range suggested by climate model
simulations of the next 100 years (6), such
warming is capable of driving changes in
the Arctic biosphere that will likely rival
any changes driven by nonclimate process-
es, including human land use. Tree growth
correlates positively with temperature in
many parts of the Arctic (Fig. 2), just as
Arctic tree growth forms and fire distur-
bance regimes change with climate (57).
Studies of recent seedling establishment
and viability north of the present treeline
(58, 59), together with evidence from an-
nually dated fossil pollen records, suggest
that Arctic plant species abundances and
ranges have varied in response to varying
temperatures (60).
The limnology of Arctic lakes will also
likely continue to respond to climate
change (61). Sediments from shallow ponds
and lakes on Ellesmere (62) (Fig. 4) and
Devon islands (63) show that warming
since the mid-19th century caused acute
shifts in diatom algal floras, including the
recent establishment in some lakes of dia-
tom populations previously too light-limit-
ed by perennial ice for their subsistence.
Freshwater algae are sensitive recorders of
environmental change and may also be har-
bingers of more profound ecological reorga-
nizations, as changes in their populations
may influence the structure of higher tro-
phic levels. The paleoenvironmental record
of the Arctic makes it clear that we should
expect changes to occur in both terrestrial
and aquatic ecosystems if the present Arctic
warming persists.
Finally, our records confirm the hypoth-
esis, originally based on a single high-Arctic
ice core record of summer temperature vari-
ations (27), that the repeated failure of Euro-
peans to find the Northwest Passage during
the 19th century was likely a result of excep-
tionally severe summertime air temperature
and sea ice extent (64). Conditions more
amenable to Arctic sea travel, and to the
discovery of the Northwest Passage, arose
naturally after the turn of the century (27).
Implications for the Future
Our reconstruction of past environmental
change in the Arctic suggests that natural
A
B
C
D
0.5
0
-0.5
-1
-1.5
0.5
0
-0.5
-1
-1.5
0.5
0
-0.5
-1
-1.5
0.5
0
-0.5
-1
-1.5 -50
0
50
100
150
1363
1364
1365
1366
1367
1368
1369
250
260
270
280
290
300
310
320
330
600
800
1000
1200
1400
1600
1600 1650 1700 1750 1800 1850 1900 1950 2000
Temperature anomaly (sigma units)
CH4 (ppbv) CO2 (ppm) Total irradiance (W/m2)SO
42- residual (ppb)
Year
Fig. 3. Comparison of
hypothesized external
climate forcing (colored
lines) and standard-
ized proxy Arctic-wide
summer-weighted annu-
al temperature [gray
lines, plotted as sigma
units (18, 19)] for (A)
atmospheric CH
4
(66),
(B) atmospheric CO
2
(67), (C) solar irradiance
(32), and (D) Greenland
(GISP2) ice core volcanic
sulfate. Eruptions known
to be overrepresented in
the GISP2 record are
marked with an asterisk
(37, 38).
SCIENCE zVOL. 278 z14 NOVEMBER 1997 zwww.sciencemag.org1254
variability is large in this region and is
working together with human forcing
(through increased concentrations of atmo-
spheric trace gases) to drive unprecedented
changes in the Arctic environment. The
complexity of natural and anthropogenic
forcing highlights the probability that as-
sumptions of climate stability, or efforts to
simply extrapolate past patterns of change
into the future, will ultimately fail to antic-
ipate future Arctic climate change and its
impacts. Reliable predictions of future
change will require climate system models
that prove effective in simulating past
changes such as those reconstructed here.
Even as these models are being developed
and tested, however, the Arctic environ-
ment is likely to continue its pace of
change.
REFERENCES AND NOTES
___________________________
1. J. T. Houghton, G. J. Jenkins, J. J. Ephraums, Eds.,
Climate Change: The IPCC Scientific Assessment
(Cambridge Univ. Press, Cambridge, 1990).
2. W. L. Chapman and J. E. Walsh, Bull. Am. Meteorol.
Soc.74, 33 (1993).
3. J. A. Curry et al., J. Clim. 9, 1731 (1996).
4. G. Weller et al., J. Biogeogr. 22, 365 (1995).
5. B. Dickson, Nature 386, 649 (1997).
6. J. T. Houghton et al., Eds., Climate Change 1995
(Cambridge Univ. Press, Cambridge, 1996).
7. P. Y. Groisman et al., Science 263, 198 (1994); P. Y.
Groisman et al., J. Clim. 7, 1633 (1994); M. C. Ser-
reze, J. A. Maslanik, J. R. Key, R. F. Kokaly, Geo-
phys. Res. Lett. 22, 2183 (1995).
8. G. C. Jacoby and R. D. D’Arrigo, Clim. Change 14,
39 (1989).
9. R. D. D’Arrigo and G. C. Jacoby, in Climate Since
A.D. 1500, R. S. Bradley and P. D. Jones, Eds.
(Routledge, London, 1992), pp. 296–311; R. D.
D’Arrigo and G. C. Jacoby, Clim. Change 25, 163
(1993).
10. R. S. Bradley and P. D. Jones, The Holocene 3, 367
(1993).
11. R. S. Bradley, Quat. Sci. Rev.9, 365 (1990).
12. L. D. Williams and R. S. Bradley, in Quaternary Envi-
ronments—Eastern Canadian Arctic, Baffin Bay and
Western Greenland, J. T. Andrews, Ed. (Allen and
Unwin, Boston, 1985), pp. 741–772.
13. H. F. Diaz et al., Arct. Alp. Res. 21, 45 (1989); J. I.
Svendsen and J. Mangerud, Clim. Dyn. 6, 213
(1992); G. M. MacDonald, T. W. D. Edwards, K. A.
Moser, R. Pienitz, J. P. Smol, Nature 361, 243
(1993); H. E. Wright Jr. et al., Eds., Global Climates
Since the Last Glacial Maximum (Univ. of Minnesota
Press, Minneapolis, 1993); P. M. Anderson and L. B.
Brubaker, Quat. Sci. Rev.13, 71 (1994); D. A. Fisher
et al., The Holocene 5, 19 (1995); S. C. Zoltai, Geogr.
Phys. Quat.49, 45 (1995); K. M. Williams et al., ibid.,
p. 13; R. B. Alley and S. Anandakrishnan, Ann. Gla-
ciol. 21, 64 (1995); P. J. H. Richard, Geogr. Phys.
Quat.49, 117 (1995); A. S. Dyke et al., ibid.50, 125
(1996).
14. J. M. Grove, The Little Ice Age (Methuen, London,
1988).
15. P. D. Jones and R. S. Bradley, in Climate Since A.D.
1500, R. S. Bradley and P. D. Jones, Eds. (Rout-
ledge, London, 1992), pp. 649–666.
16. D. Meese et al., Science 266, 1680 (1994).
17. S. R. O’Brien et al., ibid. 270, 1962 (1995).
18. The proxy temperature time series in Fig. 2 were
chosen because they were available in time series
form. Most of the paleoclimate time series have been
calibrated against instrumental data, but the scarcity
of weather stations in the Arctic makes it necessary
to rely on similar records that have been calibrated
directly against instrumental data, or to use theory, in
order to define the nature of the climate signal rep-
resented in a few of the proxy records. For example,
ice cores are commonly far from any instrumental
station, but theory supports the interpretation of ice
core melt layers as a proxy of summer temperature
[whereas ice core stable isotopic variations are more
complex on decadal time scales (10)]. Likewise, the
selection of site characteristics ensures that tree ring
chronologies correlate well with summer or annual
temperatures where they can be compared with
nearby instrumental records, reinforcing the likeli-
hood that tree ring width also correlates well with
summer or annual temperature at sites remote from
any instrumental station (8, 9, 26). Theory and (in
cases where nearby meteorological data are avail-
able) direct correlation with instrumental data sup-
port the assertion that the lake sediment records
included in this study are also useful centuries-long
proxies for early summer temperatures (20, 21, 65)
[D. R. Hardy et al.,J. Paleolimnol. 16, 227 (1996)].
Thus, although most of the time series summarized
here have been calibrated quantitatively against
summer temperature, it is more appropriate to plot
raw proxy data (x) in terms of “sigma units” defined
as normalized deviations (z) from the 1901–1960
mean (m): z5(xm)/s, where sis the standard
deviation of the raw series for the period 1901–1960.
The reference period was selected to highlight the
20th century in the context of preceding centuries.
Five-year averages are plotted to emphasize the low-
er frequency patterns of variability and to accommo-
date the ;1% temporal error that characterizes the
sediment and ice core–based records. Tree ring
chronologies generally are accurate at annual res-
olution, but they may be biased in Alaska after 1970
by moisture stress or new anthropogenic causes
(58).
19. The 400-year Arctic average summer temperature
series (Fig. 3) was generated by first averaging the
Fig. 4. Lake sediment records
from climatically and limnologi-
cally contrasting regions of Elles-
mere Island (62, 68), each show-
ing abrupt changes in the compo-
sition of fossil diatom assemblag-
es deposited within approxi-
mately the last 150 years. These
biostratigraphic changes are un-
related to differential silica preser-
vation and represent the greatest
floristic shifts of the middle to late
Holocene. Taxonomic diversifica-
tion with greater representations
of littoral and periphytic taxa (Ato
C), increased diatom algal bio-
mass (C), and recent diatom re-
colonization (D) are all consistent
with the abrupt 19th- to 20th-
century shift toward longer sum-
mer growing seasons, reduced
lake-ice severity, and greater hab-
itat availability. The limnological
consequences of the 19th- to
20th-century warming appear to
be unprecedented in the con-
text of pre–18th century natural
variability. Except where noted,
tick marks on horizontal axes
represent 10% relative frequen-
cy intervals.
ARTICLE
www.sciencemag.org zSCIENCE zVOL. 278 z14 NOVEMBER 1997 1255
standardized 5-year–averaged series (Fig. 2) for
each of the groups circled in Fig. 1 (to eliminate site
density bias) and then averaging the group averages
with the other single sites not circled in Fig. 1. Sites 8
and 29 were excluded because of their subdecadal
resolution.
20. K. A. Hughen et al., in preparation.
21. S. F. Lamoureux and R. S. Bradley, J. Paleolimnol.
16, 239 (1996).
22. K. R. Briffa et al., in Climatic Variations and Forcing
Mechanisms of the Last 2000 Years, P. D. Jones,
R. S. Bradley, J. Jouzel, Eds. (Springer-Verlag, Ber-
lin, 1996), pp. 9–41.
23. H. Garfinkel and L. B. Brubaker, Nature 286, 872
(1980).
24. Borehole climate reconstructions do not have the
high temporal resolution of the records shown in Fig.
2, but they provide a clear confirmation of the recent
warming trend seen in the annually dated proxies, as
well as useful quantitative estimates of the magni-
tude of warming [A. H. Lachenbruch et al., J. Geo-
phys. Res.87, 9301 (1982); A. H. Lachenbruch and
B. V. Marshall, Science 234, 689 (1986); S. Kakuta,
Palaeogeogr. Palaeoclimatol. Palaeoecol. 98, 225
(1992); H. Beltrami and A. E. Taylor, Global Planet.
Change 11, 127 (1995)].
25. Other quantitative records include: P. E. Calkin,
Quat. Sci. Rev.7, 159 (1988); iiii and G. C.
Wiles, in International Conference on the Role of Po-
lar Regions in Global Change, G. Weller, Ed. (Univer-
sity of Alaska, Fairbanks, AK, 1991), pp. 617–625;
C. Caseldine and J. Sto¨tter, The Holocene 3, 357
(1993); J. M. Szeicz and G. M. MacDonald, Quat.
Res.44, 257 (1995); L. H. Evison et al., The Holo-
cene 6, 17 (1996).
26. G. C. Jacoby et al., Science 273, 771 (1996).
27. R. M. Koerner, ibid. 196, 15 (1977); iiiiand D. A.
Fisher, Nature 343, 630 (1990).
28. K. R. Briffa et al., Clim. Dyn. 7, 111 (1992).
29. A. E. Jennings and N. J. Weiner, The Holocene 6,
179 (1996).
30. W. Dansgaard et al., Nature 255, 24 (1975); A. E. J.
Ogilvie, Acta Archaeol. 61, 233 (1990).
31. M. K. Hughes and H. F. Diaz, Clim. Change 26, 109
(1994).
32. J. Lean et al., Geophys. Res. Lett. 22, 3195 (1995).
33. T. J. Crowley and K.-Y. Kim, ibid. 23, 359 (1996).
34. In general, the reconstructed record of total solar
irradiance (32) agrees well with independent records
of solar activity inferred from
14
C and
10
Be cosmo-
genic isotopes (35). All three records agree that the
lowest total solar irradiance of the last 400 years
occurred in the interval 1650–1710. This period was
characterized by Arctic temperatures that, although
lower than at present, were not anomalously cold
relative to the interval 1600–1650.
35. M. Stuiver and T. F. Braziunas, The Holocene 3, 289
(1993); J. Beer et al., in The Sun as a Variable Star,
J. M. Pap, C. Fro¨lich, H. S. Hudson, S. K. Solanki,
Eds. (Cambridge Univ. Press, Cambridge, 1994),
pp. 291–300.
36. Lean et al.(32) carried out a correlation analysis
between reconstructed solar irradiance (the same
used in our Fig. 3) and the Northern Hemisphere
temperature reconstruction of Bradley and Jones
(10). For the period 1610 to 1800, they found the
solar temperature correlation to be 0.86. Assuming
that the period 1610 to 1800 was influenced by nat-
ural climate forcing only, they used regression to
define the solar-temperature relation and extrapolate
it to the present, with the resulting inference that up
to 50% of the Northern Hemisphere warming be-
tween 1860 and 1970 is attributable to solar forcing.
When we repeated the analysis using Arctic temper-
ature, we found that the 1610 to 1800 correlation
between solar irradiance and Arctic temperature
(Fig. 3) is less high at 0.47, whereas the correlation
between the volcanic activity (38) and Arctic temper-
ature (Fig. 3) over the same period is higher (–0.21,
calculated without any lags) than was suggested by
the comparison with the Northern Hemisphere
record (32). Extrapolation of the 1610 to 1800 rela-
tions to the present suggests that increasing atmo-
spheric trace-gas concentrations have been the
dominant influence on Arctic temperatures after
1920, but that solar and volcanic forcing also play a
role.
37. G. A. Zielinski et al., Science 264, 948 (1994).
38. Almost all sulfate peaks in recent polar ice core
records [in this case, the Greenland GISP2 ice core
volcanic sulfate residuals (37)] correspond to known
volcanic eruptions, and the few that are not known
from the historical record (such as the large event in
1809) are known to be real because they can also be
seen in other ice core records, including some
events recorded in the Southern Hemisphere [E.
Mosley-Thompson et al., Quat. Sci. Rev.12, 419
(1993)]. All major peaks in the GISP2 ice core sulfate
record, except the peak in 1912, correspond to a
cool event in our proxy summer temperature record
for the Arctic (Fig. 3). However, the sulfate peaks
associated with the Icelandic Laki (1783) and Alas-
kan Katmai/Novarupta (1912) eruptions, both rela-
tively proximal to Greenland, have been shown to be
overestimates of the true stratospheric volcanic op-
tical depth loading generated by these eruptions (39,
44).
39. G. A. Zielinski, J. Geophys. Res.100, 20937 (1995).
40. S. Self, M. R. Rampino, J. J. Barbera, J. Volcanol.
Geotherm. Res. 11, 41 (1981); W. B. Lyons et al.,
Ann. Glaciol. 14, 176 (1990).
41. K. R. Briffa et al., J. Geophys. Res.99, 25,835
(1994).
42. S. C. Porter, Quat. Res. 26, 27 (1986).
43. R. S. Bradley and P. D. Jones, in Climate Since A.D.
1500, R. S. Bradley and P. D. Jones, Eds. (Rout-
ledge, London, 1992), pp. 606–622; M. Sato, J. E.
Hansen, M. P. McCormick, J. B. Pollack, J. Geo-
phys. Res. 98, 22987 (1993); A. Robock and M. P.
Free, ibid.100, 11549 (1995).
44. R. B. Stothers, J. Geophys. Res. 101, 3901 (1996).
45. T. F. Stocker and L. A. Mysak, Clim. Change 20, 227
(1992).
46. D. Rind and J. T. Overpeck, Quat. Sci. Rev.12, 357
(1993).
47. G. Bond et al., Nature 365 143, (1993).
48. H. van Loon and J. C. Rogers, Mon. Weather Rev.
106, 296 (1978); J. C. Rogers and H. van Loon, ibid.
107, 509 (1979); J. W. Hurrell, Science 269, 676
(1995).
49. As shown by Jennings and Weiner (29), the Nansen
Fjord record covaries closely with the record of
Icelandic sea ice severity as reconstructed from
historical accounts. These records also show
strong covariation with the one other available cen-
turies-long decadal-resolution North Atlantic SST
record—that of L. D. Keigwin [Science 274, 1504
(1996)]. Thus, we consider the Nansen record to be
an appropriate proxy for the time-dependent vari-
ations in high-latitude North Atlantic SST over the
last 400 years.
50. P. A. Mayewski et al., Science 232, 975 (1986); P. A.
Mayewski et al., Nature 346, 554 (1990); P. A.
Mayewski et al., Atmos. Environ. 27, 2915 (1993).
51. J. A. Dowdeswell, Quat. Res.48, 1 (1997).
52. G. H. Miller, ibid.3, 561 (1973); P. T. Davis, in Qua-
ternary Environments—Eastern Canadian Arctic,
Baffin Bay and Western Greenland, J. T. Andrews,
Ed. (Allen & Unwin, Boston, 1985), pp. 682–718.
53. A. Weidick et al., Glacial Inventory and Atlas of West
Greenland, Grøndlands Geologiske Undersøgelse
Repport 158 (Greenland Geological Survey, Copen-
hagen, 1992); C. Caseldine and J. Sto¨ tter, The Ho-
locene 3, 357 (1993); A. Werner, ibid., p. 128.
54. G. H. Miller and A. de Vernal, Nature 355, 244
(1992).
55. S. L. Thompson and D. Pollard, J. Clim. 10, 871
(1997).
56. F. S. Chapin III et al., Ecology 76, 694 (1995).
57. S. Payette, L. Filion, A. Delwaide, C. Be´ gin, Nature
341, 429 (1989); Y. Bergeron and S. Archambault,
The Holocene 3, 255 (1993).
58. G. C. Jacoby and R. D. D’Arrigo, Global Biogeo-
chem. Cycles 9, 227 (1995).
59. D. J. Cooper, Arctic 39, 247 (1986).
60. J. T. Overpeck, in Climatic Variations and Forcing
Mechanisms of the Last 2000 Years, P. D. Jones,
R. S. Bradley, J. Jouzel, Eds. (Springer-Verlag, Ber-
lin, 1996), pp. 479–497.
61. D. R. Hardy and R. S. Bradley, Geosci. Can. 23, 217
(1997).
62. M. S. V. Douglas, J. P. Smol, W. Blake Jr., Science
226, 416 (1994); N. C. Doubleday, M. S. V. Douglas,
J. P. Smol, Sci. Total Environ. 160–161, 661 (1995);
A. P. Wolfe, Geol. Surv. Can. Mem., in press.
63. K. Gajewski, P. B. Hamilton, R. McNeely, J. Pale-
olimnol. 17, 215 (1997).
64. B. T. Alt, R. M. Koerner, D. A. Fisher, J. C. Bourgeois,
in The Franklin Era in Canadian Arctic History,P.D.
Sutherland, Ed. (National Museum of Canada, Otta-
wa, 1985), pp. 69–92; M. Dunbar, in Climatic
Change in Canada 5, Syllogeus 55, C. R. Harington,
Ed. (National Museum of Canada, Ottawa, 1985),
pp. 107–119; J. Newell, in Climatic Change in Can-
ada 3, Syllogeus 49, C. R. Harington, Ed. (National
Museum of Canada, Ottawa, 1983), pp. 108–129.
65. A. N. Lasca, thesis, Bates College, Lewiston, ME
(1997).
66. T. Nikazawa et al,Geophys. Res. Lett. 20, 943
(1993).
67. D. M. Etheridge et al., J. Geophys. Res.101, 4115
(1996).
68. The inflection between unsupported and back-
ground (in situ production)
210
Pb activities in the Col
Pond and Elison Lake cores indicates a date of about
1850, which coincides with the onset of major dia-
tom shifts in these cores. The Solstice Lake chronol-
ogy is based on a linear interpolation between calen-
dar age–calibrated radiocarbon ages. This model
places the major diatom changes within the last 120
years. Although the Lower Dumbell Lake core has no
radiometric chronology, it also suggests that sub-
stantial floristic change occurred within the last 100
to 150 years.
69. A. Tarussov, in Climate Since A.D. 1500, R. S. Brad-
ley and P. D. Jones, Eds. (Routledge, London,
1992), pp. 505–516.
70. K. R. Briffa et al., Nature 376, 156 (1995).
71. C. J. Earle et al., Arct. Alp. Res. 26, 60 (1994).
72. R. D. D’Arrigo et al., Can. J. For. Res.26, 143 (1996).
73. T. Kameda et al., in Proceedings of the International
Symposium on the Little Ice Age Climate, T. Mikami,
Ed. ( Tokyo Metropolitan University, Tokyo, 1992), p.
101.
74. This work is a Paleoclimates of Arctic Lakes and
Estuaries (PALE) contribution to the NSF Arctic
System Science (ARCSS) Program. It is also a
product of the NSF- and NOAA-sponsored Analy-
sis of Rapid and Recent Climatic Change (ARRCC)
and other non-PALE (for example, dendroclimatic)
projects, and is also a result of research funded by
the Natural Sciences and Engineering Research
Council of Canada (NSERC), including the Paleo-
ecological Analysis of Circumarctic Treeline
(PACT ) program. The Polar Continental Shelf
Project provided logistical field support. We thank
our colleagues for discussion of our work at
ARCSS workshops, as well as J. Cole, J. Lean, A.
Ogilvie, and two anonymous reviewers for critical
comments on the manuscript.-
SCIENCE zVOL. 278 z14 NOVEMBER 1997 zwww.sciencemag.org1256
... Available palaeoclimatological data indicates generally cold and wet conditions in Arctic North America between 1250 and 1840 CE (Overpeck et al., 1997;PAGES 2k Consortium, 2013;Linderholm et al., 2018). In agreement with peatland moisture reconstruction in Arctic Alaska/ Canada (Taylor et al., 2019b;Sim et al., 2021), our testate amoebae record and Fe/Mn ratios indicates stable, (increasingly) wet hydrological conditions with minor, at GaII potentially stronger, fluctuations in reconstructed water table depth (WTD 10 and 18 cm below peatland surface). ...
... Palaeoclimatological data indicate that there has been a temperature increase in the Arctic since ca. 1840 CE (Overpeck et al., 1997;PAGES 2k Consortium, 2013). Our palaeoecological records allow the reconstruction of past local and regional environmental changes with an unpreceded temporal resolution of 3-4 year/sample (site GaI). ...
Article
Rapidly increasing temperatures in high-latitude regions are causing major changes in wetland ecosystems. To assess the impact of concomitant hydroclimatic fluctuations, mineral deposition, and autogenous succession on the rate and direction of changing arctic plant communities in Arctic Alaska, we conducted detailed palaeoecological analyses using plant macrofossil, pollen, testate amoebae, elemental analyses, and radiocarbon and lead (210Pb) dating on two replicate monoliths from a peatland that developed in a river valley on the northern foothills of the Books Range. We observed an expansion of Sphagnum populations and vascular plants preferring dry habitats, such as Sphagnum warnstorfii, Sphagnum teres/squarrosum, Polytrichum strictum, Aulacomnium palustre and Salix sp., in recent decades between 2000 and 2015 CE, triggered by an increase in temperature and deepening water tables. Deepening peatland water tables became accentuated over the last two decades, when it reached its lowest point in the last 700 years. Conversely, a higher water-table between ca. 1500 and 1950 CE led to a recession of Sphagnum communities and an expansion of sedges. The almost continuous supply of mineral matter during this time led to a dominance of minerotrophic plant communities, although with varying species composition throughout the study period. The replicate cores show similar patterns, but nuanced differences are also visible, depicting fine spatial scale differences particularly in peat-forming plant distribution and the different timings of their presence. In conclusion, our study provides valuable insights into the impact of hydroclimatic fluctuations on peatland vegetation in Arctic Alaska, highlighting their tendency to dry out in recent decades. It also highlights the importance of river valley peatlands in paleoenvironmental reconstructions.
... Wilmking et al [88], suggest that, depending on the precipitation regime and the age of the forest, increasing temperature could positively or negatively influence tree growth. Eastern Siberia being a high-latitude region is subject to unprecedented rising temperatures due to climate change [89][90][91][92]. Our GLM results suggest that when mean annual temperatures increase in the taiga-tundra ecotone of Chukotka, where the transition zone is diffuse [93], less Sparse Larch is expected, but there is a higher likelihood of Dense Larch. ...
Article
Full-text available
The Siberian boreal forest is the largest continuous forest region on Earth and plays a crucial role in regulating global climate. However, the distribution and environmental processes behind this ecosystem are still not well understood. Here, we first develop Sentinel-2-based classified maps to show forest-type distribution in five regions along a southwest-northeast transect in eastern Siberia. Then, we constrain the environmental factors of the forest-type distribution based on a multivariate analysis of bioclimatic variables, topography, and ground-surface temperatur at the local and regional scales. Furthermore, we identify potential versus realized forest-type niches and their applicability to other sites. Our results show that mean annual temperature and mean summer and winter temperatures are the most influential predictors of forest-type distribution. Furthermore, we show that topography, specifically slope, provides an additional but smaller impact at the local scale. We find that the filling of climatic environmental niches by forest types decreases with geographic distance, but that the filling of topographic niches varies from one site to another. Our findings suggest that boreal forests in eastern Siberia are driven by current climate and topographical factors, but that there remains a portion of the variability that cannot be fully accounted for by these factors alone. While we hypothesize that this unexplained variance may be linked to legacies of the Late Glacial, further evidence is needed to substantiate this claim. Such results are crucial to understanding and predicting the response of boreal forests to ongoing climate change and rising temperatures.
... The enhanced warming rates of the Arctic regions over the past decades (ACIA, 2004;Bekryaev et al., 2010;Chapin et al., 2005;Overpeck et al., 1997;Serreze et al., 2000) have stimulated active research on the permafrost dynamics (e.g., Jorgenson et al., 2006;Oelke et al., 2003;V. E. Romanovsky et al., 2010;Yi et al., 2018). ...
Article
Full-text available
Plain Language Summary The seasonally thawed layer on top of the permafrost (active layer) is a key component of the Arctic system affected by the strong warming trend over the past decades. This soil layer experiences a pronounced seasonal cycle of freezing and thawing processes caused by the availability of Sun's energy. Mathematical modeling of the thaw depth of the active layer has remained challenging. This study formulates a novel model for the simulation of the diurnal cycle of thawing process. The formulation is developed using innovative models of heat flux that goes into the soil and soil temperature profile. Ground heat flux is derived from available energy at the land surface using a theory of surface heat flux partition. The soil temperature profile is expressed using ground heat flux within the active layer. The proposed model has been validated against field observations during thawing season. The model simulation and field observations of the thaw depth are in a good agreement at three Arctic study sites with forest and tundra surface conditions. The proposed formulation can be used for modeling freeze‐thaw cycles of the active layer at the regional scales since data on surface available energy can be obtained from remote sensing observations.
... In North American peatlands, including the lowland Tanana Flats region in Interior Alaska (263,964 ha), permafrost attained its maximum extent during the Little Ice Age (Overpeck et al., 1997;Treat & Jones, 2018;Zoltai, 1993). During periods with a relatively warm climate, including the mid-1700s and 1900s as well as the most recent decades, permafrost degradation occurs in this region, resulting in landscape transition (Jorgenson et al., 2001). ...
Article
Permafrost degradation in peatlands is altering vegetation and soil properties and impacting net carbon storage. We studied four adjacent sites in Alaska with varied permafrost regimes, including a black spruce forest on a peat plateau with permafrost, two collapse scar bogs of different ages formed following thermokarst, and a rich fen without permafrost. Measurements included year‐round eddy covariance estimates of net carbon dioxide (CO 2 ), mid‐April to October methane (CH 4 ) emissions, and environmental variables. From 2011 to 2022, annual rainfall was above the historical average, snow water equivalent increased, and snow‐season duration shortened due to later snow return. Seasonally thawed active layer depths also increased. During this period, all ecosystems acted as slight annual sources of CO 2 (13–59 g C m ⁻² year ⁻¹ ) and stronger sources of CH 4 (11–14 g CH 4 m ⁻² from ~April to October). The interannual variability of net ecosystem exchange was high, approximately ±100 g C m ⁻² year ⁻¹ , or twice what has been previously reported across other boreal sites. Net CO 2 release was positively related to increased summer rainfall and winter snow water equivalent and later snow return. Controls over CH 4 emissions were related to increased soil moisture and inundation status. The dominant emitter of carbon was the rich fen, which, in addition to being a source of CO 2 , was also the largest CH 4 emitter. These results suggest that the future carbon‐source strength of boreal lowlands in Interior Alaska may be determined by the area occupied by minerotrophic fens, which are expected to become more abundant as permafrost thaw increases hydrologic connectivity. Since our measurements occur within close proximity of each other (≤1 km ² ), this study also has implications for the spatial scale and data used in benchmarking carbon cycle models and emphasizes the necessity of long‐term measurements to identify carbon cycle process changes in a warming climate.
... The cryospheric component of Earth's biosphere has attained global research focus, as the increasing pace of glacial melts seriously alter the structural composition of the ecosystem, thereby resulting in significant change in the biogeochemical cycles and energy flow within the system [1][2][3][4][5][6][7]. Since the polar regions are considered one of the "biological hotspots," it is essential to understand the diversity, distribution and functional roles of its structural components mainly the unseen microbial drivers. ...
Article
The frosty polar environment houses diverse habitats mostly driven by psychrophilic and psychrotolerant microbes. Along with traditional cultivation methods, next-generation sequencing technologies have become common for exploring microbial communities from various extreme environments. Investigations on glaciers, ice sheets, ponds, lakes, etc. have revealed the existence of numerous microorganisms while details of microbial communities in the Arctic fjords remain incomplete. The current study focuses on understanding the bacterial diversity in two Arctic fjord sediments employing the 16S rRNA gene metabarcoding and its comparison with previous studies from various Arctic habitats. The study revealed that Proteobacteria was the dominant phylum from both the fjord samples followed by Bacteroidetes, Planctomycetes, Firmicutes, Actinobacteria, Cyanobacteria, Chloroflexi and Chlamydiae. A significant proportion of unclassified reads derived from bacteria was also detected. Psychrobacter, Pseudomonas, Acinetobacter, Aeromonas, Photobacterium, Flavobacterium, Gramella and Shewanella were the major genera in both the fjord sediments. The above findings were confirmed by the comparative analysis of fjord metadata with the previously reported (secondary metadata) Arctic samples. This study demonstrated the potential of 16S rRNA gene metabarcoding in resolving bacterial composition and diversity thereby providing new in situ insights into Arctic fjord systems.
Article
Full-text available
Unicellular eukaryotic plankton communities (protists) are the major basis of the marine food web. The spring bloom is especially important, because of its high biomass. However, it is poorly described how the protist community composition in Arctic surface waters develops from winter to spring. We show that mixotrophic and parasitic organisms are prominent in the dark winter period. The transition period toward the spring bloom event was characterized by a high relative abundance of mixotrophic dinoflagellates, while centric diatoms and the haptophyte Phaeocystis pouchetii dominated the successive phototrophic spring bloom event during the study. The data shows a continuous community shift from winter to spring, and not just a dormant spring community waiting for the right environmental conditions. The spring bloom initiation commenced while sea ice was still scattering and absorbing the sunlight, inhibiting its penetration into the water column. The initial increase in fluorescence was detected relatively deep in the water column at ~55 m depth at the halocline, at which the photosynthetic cells accumulated, while a thick layer of snow and sea ice was still obstructing sunlight penetration of the surface water. This suggests that water column stratification and a complex interplay of abiotic factors eventually promote the spring bloom initiation.
Article
Periods of formation of different sedimentary lithotypes were determined using the results of the study of modern sedimentation processes in the Storefjorden Strait (West Spitsbergen). The chronology of sedimentation was established from variations in specific activity of 210Pb and 226Ra and verified by 137Cs data. Sedimentation rates have varied considerably over the last century in different parts of the sedimentation basin, ranging from 0.04 to 0.3 cm/yr. Sedimentation in the Storefjorden Strait was closely related to changing climatic conditions. Climatic factors (air temperature, atmospheric precipitation) have a different effect in morphologically different sections of the sedimentation basin. In the south of the strait, lithotypes are determined by an oceanological factor (bottom current). Despite the changes in sedimentation rate, the lithotype in the south of Storefjorden has remained stable for more than 100 years. In the north and in the central part of the strait, cooling in the period from 1970 to 1995 caused a change in the bottom sediment structure.
Article
Full-text available
Atmospheric sounding using the Tu-134 Optik aircraft-laboratory was conducted in September 2020 over the seas of the Russian sector of the Arctic Ocean, namely the Barents, Kara, Laptev, East Siberian, Chukchi and Bering seas. Unique samples of atmospheric aerosols at altitudes from 200 and up to 10,000 m were taken, including samples for the identification of cultivated microorganisms and their genetic analysis. Data on the concentration and diversity of bacteria and fungi isolated from 24 samples of atmospheric aerosols are presented; the main phenotypic and genomic characteristics were obtained for 152 bacterial cultures; and taxonomic belonging was determined. The concentration of cultured microorganisms detected in aerosols of different locations was similar, averaging 5.5 × 103 CFU/m3. No dependence of the number of isolated microorganisms on the height and location of aerosol sampling was observed. The presence of pathogenic and condto shitionally pathogenic bacteria, including those referred to in the genera Staphylococcus, Kocuria, Rothia, Comamonas, Brevundimonas, Acinetobacter, and others, as well as fungi represented by the widely spread genera Aureobasidium, Aspergillus, Alternaria, Penicillium, capable of causing infectious and allergic diseases were present in most analyzed samples. Obtained data reveal the necessity of systematic studies of atmospheric microbiota composition to combat emerging population diseases.
Article
Full-text available
This paper presents the results of geomorphological, geochronological and lithological investigation of a young sedimentation basin – the periglacial Lake Bretjorna (Ledovoe). Formation of the lake began at the end of the first half of the 20th century in the west of Nordenskiöld Land (Western Svalbard) within the marginal zone of the Grönfjord glacier, after significant degradation of this glacial massif. The present-day look of the lake was formed at the end of the 20th century however its geomorphological development went on until the beginning of the 21st. The filling of the sedimentation basin with limnoglacial sediments began proximately from the 1930s and proceeded with spatial and temporal irregularity. Bottom sediments of the lake presented mainly by silty-pelitic material, but in places with an admixture of gravel and pebbles, were formed as the front of the Grenfjord glacier retreated in the direction from north-east to south-west. At the first stage, sediment-genesis was mainly accomplished by the intra- and subglacial sediments, which entered the lake as a result of thermal erosion at the contact of the lake and ice since the late 1940s until the end of the last century, as the present-day lake basin became ice-free. Sediments formed at this stage are composed by coarser and poorly sorted material, which is typical for glacial deposits. At the second stage, the sediments became less coarse and more sorted. After the loss of contact between the lake and the glacier, the role of intra- and subglacial sediments decreased. At this time, sedimentation in the lake goes on by the runoff of melted glacial waters and atmospheric precipitation falling on the catchment area of the lake. The value of the average sedimentation rate in the lake was determined by means of radioisotope dating according to 210Pb and 137Cs and amounted 5.4 mm/year in the north, 12.4 mm/year in the central part, and 16.4 mm/year in the south of the lake, which is comparable with the data obtained by gravity method using sedimentological traps (12–15mm/year).
Article
Full-text available
Contemporary large-scale changes in satellite-derived snow cover were examined over the Northern Hemisphere extratropical land (NEL) areas. These areas encompass 55% of the land in the Northern Hemisphere. Snow cover (S) transient regions, the `centers of action` relative to interannual variations of snow cover, were identified for the years 1972-1992. During these years a global retreat in snow cover extent (SE) occurred in the second half of the hydrologic year (April-September). Mean annual SE has decreased by 10% (2.3 x 10(exp 6) sq km). Negative trends account for one-third to one-half of the interannual continental variance of SE. The historical influence of S on the planetary albedo and outgoing longwave radiation (OLR) is investigated. The mean annual response of the S feedback on the radiative balance (RB) is negative and suggests a largescale heat redistribution. During autumn and early winter (up to January), however, the feedback of S on the planetary RB may be positive. Only by February does the cooling effect of S (due to albedo increase) dominate the planetary warming due to reduced OLR over the S. Despite a wintertime maximum in SE, the feedback in spring has the greatest magnitude. The global retreat of spring SE should lead to a positive feedback on temperature. Based on observed records of S, changes in RB are calculated that parallel an observed increase of spring temperature during the past 20 years. The results provide a partial explanation of the significant increase in spring surface air temperature observed over the land areas of the Northern Hemisphere during the past century. The mean SE in years with an El Nino and La Nina were also evaluated. El Nino events are generally accompanied by increased SE over the NEL during the first half of the hydrological year.
Article
The effect of volcanic emission of acidic aerosols on climate is well documented. The presence of acid droplets in the stratosphere can reduce transmissivity and hence decrease surface temperatures. Since the amount and chemical composition of erupted material has important effects on regional climate, knowledge of past volcanic events is of extreme importance. Detailed glaciochemical records provide the only milieu wherein the geochemistry of paleovolcanic events can be fully documented. We present a detailed sulfate and chloride record from an ice core drilled at site 20 D, 40 km SW of Dye 3 in southern Greenland. The record spans the time period 1869–1984 with chemical analyses of approximately eight samples per year. Time series decomposition and locally weighted scatter plot smoothing techniques were used to extract long term trends from the data so that individual volcanic eruptions could be documented. A number of events identified here have been unnoticed previously and a high percentage of the major chemical signatures documenting these events is associated with large decreases in temperature in the latitudinal zone 60–90 °N. Many authors have pointed out that the amount of volcanic acids such as HCl and H2SO4 injected into the atmosphere has a very important influence on global climate, yet this volcanic input has been difficult to quantify prior to ∼1960. Our data help to alleviate this problem. These individual events can be compared to available frost tree ring data from North America, further establishing a volcanism-climatic linkage.
Article
The glacier inventory and atlas provides the localisation and description of over 5000 glacier units in western Greenland between 59° 30´ and 71° 00´N. Registration is based on natural hydrological basins, in accordance with the recommendations of the International Commission on Snow and Ice (ICSI) but modified and simplified to meet Greenland conditions. The work is divided into three parts: a description of the procedure used in compilation of the glacier inventory and the glaciological conditions, a tabulated presentation of the information and description of the individual glacier units, and a 1:300 000 map presentation showing the position and delineation of the individual glacier units.
Chapter
The application of annually-laminated (varved) sediments in paleoenvironmental research has expanded significantly with the growing need for millennia-long records of interannual to century-scale climatic variability. In cases where the annual nature of the varves can be confirmed by radiometric or other methods, the varves provide annually-resolved chronostratigraphies that can be coupled with a wide range of paleoenvironmental proxies to provide time series with accurate seasonal- to centennial- scale sample resolution. In some cases, variations in the thickness and composition of the varves themselves may provide a useful paleoenvironmental signal. More commonly, a varve-based chronology is coupled with geochemical or paleontological measurements. Proper interpretations require that the climatic response time of the signal be taken into account. New and published data from eastern North America suggest that changes in relative fossil pollen abundance must lag climatic changes by a significant number of years, and that this lag is likely to be longer in forested regions than in regions dominated by non-arboreal vegetation. Vegetation model results illustrate the non-climatic factors (i.e., succession and disturbance) that can influence this lag. The climatic response times of some fossil plankton or geochemical variations are likely to involve less of a temporal lag, but the nature of the lag should be documented nonetheless. Paleoenvironmental interpretations based on multiple proxies and sites are most secure. One such interpretation suggests that the eastern Canadian Arctic experienced a dramatic warming over the last 150 years.
Article
Fluxes of trace gases from northern ecosystems represent a highly uncertain and potentially significant component of the arctic land-atmosphere system, especially in the context of greenhouse-induced climate change. The initial goal of the Arctic Flux Study (a part of NSF's Arctic System Science Program) is a regional estimate of the present and future movement of materials between the land, atmosphere and ocean in the Kuparuk River basin in northern Alaska. We are measuring rates and controls of processes along a north-south transect running from the marshy coastal plain to mountain valleys. Important vertical fluxes under study are the release of CO2 and CH4 from soils and water, lateral fluxes are surface water, nutrients, and organic matter. A hierarchy of measurements allow the rates and understanding of processes to be scaled from plots to the landscape, regional, and circumarctic level. These include gas flux measurements in small chambers, measurements over larger areas by eddy correlation from small towers, and measurements at the landscape scale from airplane overflights. Experimental manipulations of carbon dioxide, soil moisture, nutrients and soil temperature from this and other studies give information on process controls. The distribution of plant communities has been described at several landscape-scale sites and a hierarchic GIS has been developed for the region at three scales (plot, landscape, region). Climate is measured at six sites and hydrological processes are being studied at each watershed scale. In the soils, measurements are being made of soil organic matter and active layer thickness and of availability of soil organic matter for microbial transformation into CO2 and CH4. Fluxes and process understanding have been incorporated into a hierarchy of models at different scales. These include models of regional climate nested in a GCM; of regional-and continental-scale plant productivity and carbon cycling incluing CO2 release under altered climates; watershed and regional models of hydrology; and surface energy budgets. After the first year of study the regional climate model has been successfully configured to the northern Alaska region. We have also measured a large release of carbon dioxide from tundra soils in all but the coldest and wettest parts of the transect. The rates from eddy correlation towers (landscape level) agree closely with rates from chambers (plot level). Observations, experimental manipulations and modelling analyses result in the prediction that the combination of warmer and drier soils is responsible for the large CO2 release.
Article
To provide a background for ARM`s activities at the North Slope of Alaska/Adjacent Arctic Ocean sites, an overview is given of our current state of knowledge of Arctic cloud and radiation properties and processes. The authors describe the Arctic temperature and humidity characteristics, cloud properties and processes, radiative characteristics of the atmosphere and surface, direct and indirect radiative effects of aerosols, and the modeling and satellite remote sensing of cloud and radiative characteristics. An assessment is given of the current performance of satellite remote sensing and climate modeling in the Arctic as related to cloud and radiation issues. Radiation-climate feedback processes are discussed, and estimates are made of the sign and magnitude of the individual feedback components. Future plans to address these issues are described. 276 refs., 12 figs.