ArticlePDF Available

Past temperature directly from the Greenland ice sheet

Authors:

Figures

Content may be subject to copyright.
DOI: 10.1126/science.282.5387.268
, 268 (1998);282 Science , et al.D. Dahl-Jensen
Past Temperatures Directly from the Greenland Ice Sheet
This copy is for your personal, non-commercial use only.
clicking here.colleagues, clients, or customers by , you can order high-quality copies for yourIf you wish to distribute this article to others
here.following the guidelines can be obtained byPermission to republish or repurpose articles or portions of articles
): February 3, 2011 www.sciencemag.org (this infomation is current as of
The following resources related to this article are available online at
http://www.sciencemag.org/content/282/5387/268.full.html
version of this article at: including high-resolution figures, can be found in the onlineUpdated information and services,
http://www.sciencemag.org/content/282/5387/268.full.html#ref-list-1
, 2 of which can be accessed free:cites 20 articlesThis article
339 article(s) on the ISI Web of Sciencecited by This article has been
http://www.sciencemag.org/content/282/5387/268.full.html#related-urls
40 articles hosted by HighWire Press; see:cited by This article has been
http://www.sciencemag.org/cgi/collection/geochem_phys
Geochemistry, Geophysics subject collections:This article appears in the following
registered trademark of AAAS. is aScience1998 by the American Association for the Advancement of Science; all rights reserved. The title CopyrightAmerican Association for the Advancement of Science, 1200 New York Avenue NW, Washington, DC 20005.
(print ISSN 0036-8075; online ISSN 1095-9203) is published weekly, except the last week in December, by theScience
on February 3, 2011www.sciencemag.orgDownloaded from
nated to a surface Al, providing a mechanism
for the interchange of O
s
and O
ads
. These re-
sults also provide evidence for incipient
Al(OH)
3
formation on the surface. The ultimate
structure of the heavily hydrated surface is
clearly very complicated and may depend
strongly on sample history. It is possible that
the Al(OH)
3
species can be removed complete-
ly (perhaps starting near steps or other defects),
leaving a less reactive surface that is completely
O
s
H-terminated, which is similar to the known
surfaces of aluminum hydroxides (1).
An idealized model for fully hydroxylated
a-Al
2
O
3
(0001) (28) replaces each surface Al
with three H atoms (Fig. 5A), yielding a
coverage .15 OH per square nanometer.
Room-temperature MD simulations of this
model revealed a complex dynamic structure
(Fig. 5B), with one out of every three OH
groups, on average, lying parallel to the sur-
face because of in-plane hydrogen bonding.
Calculated O–H vibrational spectra (25)
yielded two broad peaks at ;3470 and 3650
cm
–1
, with the peak at ;3470 cm
–1
corre-
sponding to in-plane OH groups. The peak at
;3650 cm
–1
is close to the single peak (3720
to 3733 cm
–1
) that is observed in most mea-
surements on hydroxylated a-Al
2
O
3
(0001)
(29) and to the range that is generally as-
signed to bridging OH groups (2,26). The
peak at ;3470 cm
–1
is red-shifted by hydro-
gen bonding and is generally not seen in
single-crystal experiments, perhaps because
of selection rules or because it is too broad.
Our finding of two peaks split by 200 cm
–1
contradicts all previous classifications of OH
vibrations (and subsequent cluster modeling)
(2) on aluminas, which assume that all OH
groups with the same coordination of O and
neighboring Al have the same frequency. By
this criterion, all of the surface OH groups in
Fig. 5 are equivalent; however, their stretch
frequencies clearly depend also on longer
range environmental effects.
The present investigation of a-Al
2
-
O
3
(0001) has elucidated several aspects of
the complex interactions of H
2
O with an
alumina surface, especially the dynamics of
dissociation reactions at low and high cover-
ages. On the basis of these results, a consis-
tent interpretation of a diverse set of experi-
mental data on hydroxylated alumina surfac-
es begins to emerge.
References and Notes
1. K. Wefers and C. Misra, Alcoa Tech. Pap. 19 (revised)
(Alcoa Laboratories, St. Louis, MO, 1987).
2. H. Kno¨zinger and P. Ratnasamy, Catal. Rev. Sci. Eng.
17, 31 (1978).
3. M. Gautier et al.,J. Am. Ceram. Soc. 77, 323 (1994).
4. P. de Sainte Claire, K. C. Hass, W. F. Schneider, W. L.
Hase, J. Chem. Phys. 106, 7331 (1997).
5. G. N. Robinson, Q. Dai, A. Freedman, J. Phys. Chem. B
101, 4940 (1997).
6. R. Car and M. Parrinello, Phys. Rev. Lett. 55, 2471
(1985).
7. The gradient-corrected exchange-correlation [Bern-
stein, Lee, Yang, and Primakoff (BLYP)] functional
used here is from A. D. Becke [Phys. Rev. A 38, 3098
(1988)] and C. Lee, W. Yang, and R. Parr [Phys. Rev. B
37, 785 (1988)]. Norm-conserving numerical pseudo-
potentials were generated for Al and O with the
procedure of N. Troullier and J. L. Martins [ibid. 43,
1993 (1991)], and a local analytic pseudopotential
was derived for H. This is essentially a softened
Coulomb potential with a core radius of 0.25 atomic
units. Electron wave functions are expanded in a
plane-wave basis set with an energy cutoff of 70
rydbergs (Ry). We used the Car-Parrinello Molecular
Dynamics code in the parallelized 2.5 version (devel-
oped by J. Hutter and copyrighted by IBM, Armonk,
NY). All calculations were performed on a 32-node
IBM RS6000 SP at the IBM Watson Research Labora-
tory (Yorktown Heights, NY ).
8. In the MD runs, a value of 400 au was used for the
fictitious electron mass of the Car-Parrinello Lagrang-
ian multipliers (6), and each hydrogen molecule was
replaced by deuterium to improve the separation
between electronic and ionic degrees of freedom. The
time step in the Verlet algorithm for the integration
of the equations of motions was ;0.1 fs.
9. The importance of chemical reaction dynamics in
general has recently been highlighted in a special
issue of Science [Reaction Dynamics, Science 279,
1875–1895 (1998)].
10. A. Curioni et al.,J. Am. Chem. Soc. 119, 7218 (1997).
11. V. E. Puchin et al.,Surf. Sci. 370, 190 (1997); J. Ahn
and J. W. Rabalais, ibid. 388, 121 (1997).
12. See, for example, S. Blonski and S. H. Garofalini, ibid.
295, 263 (1993).
13. See, for example, M. Causa, R. Dovesi, C. Pisani, C.
Roetti, ibid. 215, 259 (1989); I. Manassidis, A. De
Vita, M. J. Gillan, Surf. Sci. Lett. 285, L517 (1993); I.
Frank, D. Marx, M. Parrinello, J. Chem. Phys. 104,
8143 (1996).
14. J. M. McHale, A. Auroux, A. J. Perrotta, A. Navrotsky,
Science 277, 788 (1997). For earlier work, see P. A.
Thiel and T. E. Madey, Surf. Sci. Rep. 7, 211 (1987)
and references therein.
15. J. M. Wittbrodt, W. L. Hase, H. B. Schlegel, J. Phys.
Chem. B 102, 6539 (1998).
16. K. C. Hass, W. F. Schneider, A. Curioni, W. Andreoni, in
preparation.
17. Earlier calculations used much smaller supercells than
the present work. Such studies were therefore limited
in their ability to provide accurate adsorbate struc-
tures and energies and to study the H
2
O coverage
dependence and phenomena such as collective ef-
fects and surface diffusion.
18. J. Goniakowski and M. J. Gillan, Surf. Sci. 350, 145
(1996); P. J. D. Lindan, N. M. Harrison, J. M. Hold-
ender, M. J. Gillan, Chem. Phys. Lett. 261, 246 (1996);
P. J. D. Lindan, N. M. Harrison, M. J. Gillan, Phys. Rev.
Lett. 80, 762 (1998).
19. W. Langel and M. Parrinello, J. Chem. Phys. 103, 3240
(1995).
20. Lagrange multipliers were introduced to constrain
the relevant H–O
s
distance, and the average con-
straint forces were determined from constant tem-
perature simulations [S. Nose´, J. Chem. Phys. 81, 511
(1984); W. G. Hoover, Phys. Rev. A 31, 1695 (1985)]
of at least 0.2 ps.
21. S. Scheiner, in Proton Transfer in Hydrogen-Bonded
Systems, T. Bountis, Ed. (Plenum, New York, 1992), p.
29.
22. S. Blonski and S. H. Garofalini, J. Phys. Chem. 100,
2201 (1996).
23. The temperature was not controlled but was in-
creased slowly from ;100 to ;350 K. The system
was then allowed to evolve for a time interval of .1
ps. The average temperature was 250 K.
24. D. E. Brown, D. J. Moffatt, R. A. Wolkow, Science 279,
542 (1998).
25. Vibrational frequencies were estimated from the
power spectra of the (partial) velocity-velocity auto-
correlation functions and were rescaled to account
for the fictitious electronic mass and the different
mass used for the proton.
26. V. I. Lygin and I. S. Muzyka, Russ. J. Phys. Chem. 69,
1829 (1995); A. Tsyganenko and P. Mardilovich, J.
Chem. Soc. Faraday Trans. 92, 4843 (1996).
27. B. A. Huggins and P. D. Ellis, J. Am. Chem. Soc. 114,
2098 (1992).
28. M. A. Nygren, D. H. Gay, C. R. A. Catlow, Surf. Sci.
380, 113 (1997).
29. C. Morterra, G. Ghiotti, E. Garrone, F. Boccuzzi, J.
Chem. Soc. Faraday Trans. 1 72, 2722 (1976); J. G.
Chen, J. E. Crowell, J. T. Yates, J. Chem. Phys. 84, 5906
(1986); V. Coustet and J. Jupille, Surf. Sci. 307, 1161
(1994).
11 August 1998; accepted 3 September 1998
Past Temperatures Directly
from the Greenland Ice Sheet
D. Dahl-Jensen,* K. Mosegaard, N. Gundestrup, G. D. Clow,
S. J. Johnsen, A. W. Hansen, N. Balling
A Monte Carlo inverse method has been used on the temperature profiles
measured down through the Greenland Ice Core Project (GRIP) borehole, at the
summit of the Greenland Ice Sheet, and the Dye 3 borehole 865 kilometers
farther south. The result is a 50,000-year-long temperature history at GRIP and
a 7000-year history at Dye 3. The Last Glacial Maximum, the Climatic Optimum,
the Medieval Warmth, the Little Ice Age, and a warm period at 1930 A.D. are
resolved from the GRIP reconstruction with the amplitudes –23 kelvin, 12.5
kelvin, 11 kelvin, –1 kelvin, and 10.5 kelvin, respectively. The Dye 3 temper-
ature is similar to the GRIP history but has an amplitude 1.5 times larger,
indicating higher climatic variability there. The calculated terrestrial heat flow
density from the GRIP inversion is 51.3 milliwatts per square meter.
Measured temperatures down through an ice
sheet relate directly to past surface tempera-
ture changes. Here, we use the measurements
from two deep boreholes on the Greenland
Ice Sheet to reconstruct past temperatures.
The GRIP ice core (72.6°N, 37.6°W) was
successfully recovered in 1992 (1, 2), and the
3028.6-m-deep liquid-filled borehole with a
diameter of 13 cm was left undisturbed. Tem-
peratures were then measured down through
the borehole in 1993, 1994, and 1995 (3, 4).
We used the measurements from 1995 (Fig.
1) (4), because there was no remaining evi-
dence of disturbances from the drilling and
REPORTS
9 OCTOBER 1998 VOL 282 SCIENCE www.sciencemag.org268
on February 3, 2011www.sciencemag.orgDownloaded from
the measurements were the most precise (65
mK). Temperatures measured in a thermally
equilibrated shallow borehole near the drill
site are used for the top 40 m, because they
are more reliable than the GRIP profile over
this depth (5). The present mean annual sur-
face temperature at the site is –31.70°C. The
2037-m-deep ice core from Dye 3 (65.2°N,
43.8°W) was recovered in 1981. We used
temperature data from 1983 measurements
with a precession of 30 mK (6, 7). The tem-
peratures at the bedrock are –8.58°C at GRIP
and –13.22°C at Dye 3. Calculations show
that the basal temperatures have been well
below the melting point throughout the past
100,000 years (8). Because there are still
climate-induced temperature changes near
the bedrock, we included 3 km of bedrock in
the heat flow calculatin.
Past surface temperature changes are in-
dicated from the shape of the temperature
profiles (Fig. 1). We used a coupled heat- and
ice-flow model to extract the climatic infor-
mation from the measured temperature pro-
files. The temperatures down through the ice
depend on the geothermal heat flow density
(heat flux), the ice-flow pattern, and the past
surface temperatures and accumulation rates.
The past surface temperatures and the geo-
thermal heat flow density are unknowns,
whereas the past accumulation rates and ice-
flow pattern are assumed to be coupled to the
temperature history through relations found
from ice-core studies (911). The total ice
thickness is assumed to vary 200 m as de-
scribed in (9). The coupled heat- and ice-flow
equation is (7, 9, 12)
rc]T
]t5¹z~K¹T!2rcnYz¹T1f
where T(x,z,t) is temperature, tis time, zis
depth, xis horizontal distance along the flow
line, r(z) is ice density, K(T,r) the thermal
conductivity, c(T) is the specific heat capac-
ity, and f(z) is the heat production term. The
ice velocities, vY(x,z,t), are calculated by an
ice-flow model (9, 13). Model calculations to
reproduce a present-day temperature profile
through the ice sheet are started 450,000
years ago (ka) at GRIP (100 ka at Dye 3),
D. Dahl-Jensen, K. Mosegaard, N. Gundestrup, S. J.
Johnsen, A. W. Hansen, Niels Bohr Institute for As-
tronomy, Physics and Geophysics, Department of
Geophysics, Juliane Maries Vej 30, DK-2100 Copen-
hagen OE, Denmark. G. D. Clow, USGS-Climate Pro-
gram, Box 25046, MS 980, Denver Federal Center,
Denver, CO 80225, USA. N. Balling, Department of
Earth Sciences, Geophysical Laboratory, University of
Aarhus, Finlandsgade 8, DK-8200 Aarhus N, Denmark.
*To whom correspondence should be addressed. E-
mail: ddj@gfy.ku.dk
020 0
1000
500
1000
1000
2000
3000 2000
1500
1000
500
0
-30
-31.8 -20 -18
-31.6 -31.4
CO CO
LIA LIA
-20 -20 -18 -16 -14-10
Temperature (°C)
A
B
C
D
Temperature (°C)
Depth (m)
Depth (m)
GRIP Temperature Profile 1995 Dye 3 Temperature Profile 1983 Fig. 1. The GRIP and Dye 3 temperature profiles
[blue trace in (A) and (C)] are compared to
temperature profiles [red trace in (A) and (C)]
calculated under the condition that the present
surface temperatures and accumulation rates
have been unchanged back in time. (A) The
GRIP temperature profile measured in 1995.
The cold temperatures from the Glacial Period
(115 to 11 ka) are seen as cold temperatures
between 1200- to 2000-m depth. (B) The top
1000 m of the GRIP temperature profiles are
enlarged so the Climatic Optimum (CO, 8 to 5
ka), the Little Ice Age (LIA, 1550 to 1850 A.D.),
and the warmth around 1930 A.D. are indicated
at the depths around 600, 140, and 60 m,
respectively. (C) The Dye 3 temperature profile
measured in 1983. Note the different shape of
the temperature profiles when compared to
GRIP and the different depth locations of the
climate events. (D) The top 1500 m of the Dye
3 temperature profiles are enlarged so the CO,
the LIA, and the warmth around 1930 A.D. are
indicated at the depths around 800, 200, and
70 m, respectively.
700
600
500
400
300
200
100
00
100
250
200
150
100
50
00
100
200
300
400
500
600
Temperature (°C)
A
DE
BCF
Heat flux (mW/m2)
Geothermal heat flux1000 A.D.1600 A.D.1970 A.D.
Number of solutionsNumber of solutions
8 ka BP 25 ka BP
200
300
400
500
0
100
200
300
400
500 450
400
350
300
250
200
150
100
50
0
-32.5 -32 -31.5 -31 -30.5 -33 -32 -31 -31.5 -31 -30.5 -30 -29.5
-29.5-30-30.5-31-31.5 50.5 51 51.5 52 52.5
-51-52-53-54-55-56-57-58-59
Fig. 2. (Athrough E) The probability distributions of the past surface temperatures at the
Greenland Ice Sheet summit at selected times before present. They are constructed as histograms
of the 2000 Monte Carlo sampled and accepted temperature histories (17). All temperature
distributions are seen to have a zone with maximum values, the most likely values, which are
assumed to be the reconstructed surface temperature at these times (18). (F) The probability
distribution of the sampled geothermal heat flow densities. The most likely value is 51.3 mW/m
2
.
REPORTS
www.sciencemag.org SCIENCE VOL 282 9 OCTOBER 1998 269
on February 3, 2011www.sciencemag.orgDownloaded from
more than twice the time scale for thermal
equilibrium of the ice-bedrock, so the un-
known initial conditions are forgotten when
generating the most recent 50,000-year tem-
perature history (7000 years for Dye 3).
We developed a Monte Carlo method to fit
the data and infer past climate. The Monte
Carlo method tests randomly selected combina-
tions of surface temperature histories and geo-
thermal heat flow densities by using them as
input to the coupled heat- and ice-flow model
and considering the resulting degrees of fit be-
tween the reproduced and measured tempera-
ture profiles (1416). Our results for each site
are based on tests of 3.3 310
6
combinations
of temperature histories and heat flow densities,
of which 2000 solutions have been selected
(17). The 2000 temperature histories and heat
flow densities are sampled with a frequency
proportional to their likelihood (14, 15), and all
accepted solutions fit the observations within
their limits of uncertainty.
Histograms of the sampled geothermal heat
flow densities and of the temperature histories
at each time before present can be made (for
example, Fig. 2). The distributions in general
show that there is a most likely value, a maxi-
mum, at all times, which we refer to as the
temperature history (18). The distribution of
accepted geothermal heat flow densities (Fig.
2F) has a median of 51.3 60.2 mW/m
2
, which
is slightly higher than the heat flow density
from Archean continental crust across the Baf-
fin Bay in Canada. A few heat flow measure-
ments have been made from the coast of Green-
land (36 and 43 mW/m
2
), but these are not
corrected for long-term climate variations and
are minimum values (19). The homogeneous
thermal structure of ice is an advantage when
the heat flow density and the temperature his-
tory are to be reconstructed (20).
Histograms from the GRIP reconstruction
(Fig. 3) show that temperatures at the Last
Glacial Maximum (LGM) were 23 62K
colder than at present (21). The temperatures
at this time, 25 ka, reflect the cold temperatures
seen on the measured temperature profile at a
depth of 1200 to 2000 m. Alternative recon-
structions of the ice thickness and accumulation
rates all reproduce LGM temperatures within 2
K(9, 10, 22, 23). The cold Younger Dryas and
the warm Bølling/Allerød periods (24) are not
resolved in the inverse reconstruction. The tem-
perature signals of these periods have been
obliterated by thermal diffusion because of their
short duration (25). After the termination of the
glacial period, temperatures in our record in-
crease steadily, reaching a period 2.5 K warmer
than present during what is referred to as the
Climatic Optimum (CO), at 8 to 5 ka. Follow-
ing the CO, temperatures cool to a minimum of
0.5 K colder than the present at around 2 ka.
The record implies that the medieval period
around 1000 A.D. was 1 K warmer than present
in Greenland. Two cold periods, at 1550 and
1850 A.D., are observed during the Little Ice
Age (LIA) with temperatures 0.5 and 0.7 K
below the present. After the LIA, temperatures
reach a maximum around 1930 A.D.; tempera-
tures have decreased during the last decades
(26). The climate history for the most recent
times is in agreement with direct measurements
in the Arctic regions (27). The climate history
for the last 500 years agrees with the general
understanding of the climate in the Arctic re-
gion (28) and can be used to verify the temper-
ature amplitudes. The results show that the
temperatures in general have decreased since
the CO and that no warming in Greenland is
observed in the most recent decades.
As seen in Fig. 3, resolution decreases back
AB
C
Past temperatures (°C)
Years before present (ka) Years before present (ka)
Years A.D.
Fig. 3. The contour plots of all the GRIP temperature histograms as a function of time describes the
reconstructed temperature history (red curve) and its uncertainty. The temperature history is the
history at the present elevation (3240 m) of the summit of the Greenland Ice Sheet (21). The white
curves are the standard deviations of the reconstruction (18). The present temperature is shown as
a horizontal blue curve. The vertical colored bars mark the selected times for which the temper-
ature histograms are shown in Fig. 2. (A) The last 100 ky BP. The LGM (25 ka) is seen to have been
23 K colder than the present temperature, and the temperatures are seen to rise directly into the
warm CO 8 to 5 ka. (B) The last 10 ky BP. The CO is 2.5 K warmer than the present temperature,
and at 5 ka the temperature slowly cools toward the cold temperatures found around 2 ka. (C) The
last 2000 years. The medieval warming (1000 A.D.) is 1 K warmer than the present temperature,
and the LIA is seen to have two minimums at 1500 and 1850 A.D. The LIA is followed by a
temperature rise culminating around 1930 A.D. Temperature cools between 1940 and 1995.
-30
-32
-34
86420
200015001000500
-31
-17.5
-20.0
-22.5
Dye 3
GRIP
-18
-19
-20
-21
Years before present (ka)
A
B
Years A.D.
Dye 3 past surface temperatures (°C)
GRIP past surface temperatures (°C)
-32
Fig. 4. The reconstructed temperature histories
for GRIP (red curves) and Dye 3 (blue curves) are
shown for the last 8 ky BP (A) and the last 2 ky BP
(B). The two histories are nearly identical, with
50% larger amplitudes at Dye 3 than found at
GRIP. The reconstructed climate must represent
events that occur over Greenland, probably the
high-latitude North Atlantic region.
REPORTS
9 OCTOBER 1998 VOL 282 SCIENCE www.sciencemag.org270
on February 3, 2011www.sciencemag.orgDownloaded from
in time (25, 29). For the GRIP reconstruction,
an event with a duration of 50 years and an
amplitude of 1 K can be resolved 150 years
back in time with a measurement accuracy of 5
mK; an event with a similar amplitude but a
duration of 1000 years can be detected back to
5 ka. An event that occurred 50 ka will now be
observed in the temperature profile at the bed-
rock. Climate events for times older than
50,000 years before present (ky BP) are not
well resolved (30). At Dye 3, the reconstructed
climate history extends only to 7 ka, because
the ice is 1000 m thinner than at the summit and
surface accumulation rate is 50% higher. The
LGM is not well resolved in the Dye 3 record,
and consequently the geothermal heat flow den-
sity is not uniquely determined (31). On the
other hand, the recent climate history has a
higher resolution because of the increased ac-
cumulation (Fig. 4).
The Dye 3 record is nearly identical with the
GRIP record back to 7 ka, but the amplitudes
are 50% higher. Thus, the resolved climate
changes have taken place on a regional scale;
many are seen throughout the Northern Hemi-
sphere (27, 28, 32). GRIP is located 865 km
north of Dye 3 and is 730 m higher in elevation.
Surface temperatures at the summit are influ-
enced by maritime air coming in from the North
Atlantic and air masses arriving from over
northeastern Canada (associated with the Baffin
trough) (28, 32, 33). Temperatures at Dye 3 will
be influenced to a greater degree by the North
Atlantic maritime air masses. Dye 3 is located
closer to the center of the highest atmospheric
variability, which is associated with large in-
terseasonal, interannual, and decadal tempera-
ture changes (32, 34). It is therefore believed
that the observed difference in amplitudes be-
tween the two sites is a result of their different
geographic location in relation to variability of
atmospheric circulation, even on the time scale
of a millennium.
References and Notes
1. Greenland Ice-Core Project (GRIP) members, Nature
364, 203 (1993).
2. W. Dansgaard et al.,ibid., p. 218.
3. N. S. Gundestup, D. Dahl-Jensen, S. J. Johnsen, A.
Rossi, Cold Reg. Sci. Technol. 21, 399 (1993).
4. G. D. Clow, R. W. Saltus, E. D. Waddington, J. Glaciol.
42, 576 (1996).
5. The deep borehole is located in a building, and the liquid
surface in the borehole is found at a depth of 40 m. The
temperatures measured in the top 40 m are very dis-
turbed, so we used measurements from an air-filled
shallow borehole (100 m) near the borehole.
6. N. S. Gundestrup and B. L. Hansen, J. Glaciol. 30, 282
(1984).
7. D. Dahl-Jensen and S. J. Johnsen, Nature 320, 250
(1986).
8. D. Dahl-Jensen et al.,J. Glaciol. 43, 300 (1997).
9. Between 50 and 20 ka, the ice thickness was 50 m
less than at present, even though the ice sheet cov-
ered a larger area. The maximum ice thickness of
3230 m is found at 10 ka, after which the ice thick-
ness gradually has decreased to the present 3028.6
m. The depression and uplift of the bedrock influenc-
es the elevation of the surface [S. J. Johnsen, D.
Dahl-Jensen, W. Dansgaard, N. S. Gundestrup, Tellus
B47, 624 (1995)].
10. K. M. Cuffey and G. D. Clow, J. Geophys. Res. 102,
26383 (1997).
11. The past accumulation rates are determined by cou-
pling them to the past (unknown) temperature
through the relation l(T)5l
O
exp[0.0467(TT
O
)–
0.000227(TT
O
)
2
], where l(T) is the accumulation
rate at the surface temperature T,l
O
is the present
ice accumulation rate, which is 0.23 m/year at GRIP
and 0.49 m/year at Dye 3, and T
O
is the present
surface temperatures at the sites: –31.7°C at GRIP
and –20.1°C at Dye 3, respectively (9).
12. S. J. Johnsen, IAHS-AISH Publ. 118, 388 (1977).
13. S. J. Johnsen and W. Dansgaard, NATO ASI Ser. I
Global Environ. Change 2, 13 (1992).
14. K. Mosegaard and A. Tarantola, J. Geophys. Res. 100,
12431 (1995).
15. K. Mosegaard, Inverse Problems 14, 405 (1998).
16. Our Monte Carlo scheme is a random walk in the high
dimensional space of all possible models, m (tempera-
ture histories and geothermal heat flow densities). The
temperature history has been divided in 125 intervals
(interval length is 25 ky at 450 ka and 10 years at
present). Including the geothermal heat flow density as
an unknown the model space is 126-dimensional. In
each step of the random walk, a perturbed model, m
pert
i
of the current model vector m
i
is proposed. The next
model becomes equal to m
pert
i
with an acceptance
probability P
accept
5min{1,exp(–[S(m
pert
i
)–S(m
i
)])},
where S(m)5S
j
(g
j
(m)–d
obs
j
)
2
, which is the misfit
function measuring the difference between g(m), the
calculated borehole temperatures, and d
obs
, the ob-
served temperatures. If the perturbated model is reject-
ed, the next model becomes equal to m
i
and a new
perturbed model is proposed. To ensure an efficient
sampling of all possible models, we developed ways of
choosing the temperature histories and geothermal
heat flow densities to be tested. The main scheme to
perturb the models is to randomly select one of the 126
temperature/heat flow density parameters and change
its value to a new value chosen uniformly at random
within a given interval. A singular value decomposition
(SVD) of the matrix G5{]g
j
/]m
i
}, evaluated in a
near-optimal model, yields a set of eigenvectors in the
model space whose orientations reveal efficient direc-
tions of perturbation for the random walk. The SVD
method is included as a possible method of perturbing
models especially in the start of the process as it speeds
the Monte Carlo scheme significantly.
17. Of the 3.3 310
6
models tested during the random
walk 30% have been accepted by the Monte Carlo
scheme (16). Every 500 is chosen of those where the
misfit function S (16) is less than the variance of the
observations. The waiting time of 500 has been cho-
sen to exceed the maximum correlation length of the
output model parameters. This is a necessary condi-
tion for the 2000 models to be uncorrelated. To
further ensure that the output models were uncor-
related, the random walk was frequently restarted at
several random selected points in the model space.
18. The probabilistic formulation of the inverse problem
leads to definition of a probability distribution in the
model space, describing the likelihood of possible
temperature histories and geothermal heat flow den-
sities. The Monte Carlo scheme is constructed to
sample according to this probability distribution. The
histograms in Fig. 2 describe the probability distribu-
tion of the geothermal heat flow density and tem-
peratures at times before present. The maxima in the
histograms thus describe the most likely values. The
method does not constrain the distributions to have
a single maximum, indeed there could be histograms
with several maxima, reflecting that more than one
value of the temperature at this time would give a
good fit to the observed temperature in the borehole.
The histograms however, are all seen to have a
well-defined zone with most likely past tempera-
tures. A soft curve is fitted to the histograms and the
maximum value is taken as the most likely value. The
standard deviations shown in Fig. 3 are derived as
deviations from the maximum value.
19. J. H. Sass, B. L. Nielsen, H. A. Wollenberg, R. J. Munroe,
J. Geophys. Res. 77, 6435 (1972).
20. C. Clauser et al.,ibid. 102, 18417 (1997); L. Guillou-
Frottier, J.-C. Marescal, J. Musset, ibid. 103, 7385
(1998); H. N. Pollack, S. J. Hurter, J. R. Johnson, Rev.
Geophys. 31, 267 (1993); W. G. Powell, D. S. Chap-
man, N. Balling, A. E. Beck, in Handbook of Terrestrial
Heat-Flow Density Determination (Kluwer Academic,
New York, 1988), pp. 167–222.
21. In order to produce a past temperature record from
the calculated past surface temperatures, the tem-
peratures have been corrected to the present eleva-
tion of the GRIP site (and Dye 3 site respectively)
using the surface elevation changes described in (9)
and a lapse rate of 0.006 K/m.
22. K. M. Cuffey et al.,Science 270, 455 (1995).
23. D. Dahl-Jensen, in Proceedings of the Interdisciplinary
Inversion Workshop 2, Copenhagen, 19 May 1993, K.
Mosegaard, Ed. (The Niels Bohr Institute for Astron-
omy, Physics and Geophysics, University of Copen-
hagen, Copenhagen, 1993), pp. 11–14.
24. C. U. Hammer et al.,Report on the stratigraphic
dating of the GRIP Ice Core.Special Report of the
Geophysical Department (Niels Bohrs Institute for
Astronomy, Physics and Geophysics, University of
Copenhagen, Copenhagen, in press).
25. J. Firestone, J. Glaciol. 41, 39 (1995).
26. The amplitude of the warming at 1930 A.D. must be
considered to be more uncertain. The information
leading to this result are the measured temperatures
in an open shallow borehole, where air movements
could influence the measurements.
27. D. Fisher et al.,NATO ASI Ser. I Global Environ.
Change 41, 297 (1996); J. W. C. White et al.,J.
Geophys. Res. 102, 26425 (1997); J. W. Hurrel, Sci-
ence 269, 676 (1995); P. Frich et al.,inDMI Scientific
Report 96-1 (Danish Meteorological Institute, Copen-
hagen, 1996).
28. R. G. Barry and R. J. Charley, Atmosphere, Weather &
Climate (Routledge, London, ed. 6, 1992); J. Overpeck
et al.,Science 278, 1251 (1997); H. H. Lamb, Climate
History and the Modern World (Routledge, London,
ed. 2, 1995); N. W. T. Brink and A. Weidick, Quat. Res.
4, 429 (1974).
29. G. D. Clow, Palaeogeogr. Palaeoclimatol. Palaeoecol.
98, 81 (1992).
30. To comply with this resolution the time steps have
been chosen with increasing length back in time. The
increasing length of the time steps can be considered
as an efficient way of calculating the mean temper-
atures in the intervals so full available resolution is
kept but the calculations are rationalized.
31. In (7), it is argued that parameter combinations of
mean glacial temperature, mean glacial accumula-
tion, and geothermal heat flow density can be found
that fit the Dye 3 measurements due to the reduced
resolution of the climate history reaching further
back than 7 ka. A combination with a geothermal
heat flow density of 38.7 mW/m
2
was chosen corre-
sponding to a mean glacial temperature 12 K colder
than the present temperatures. If a value of 51
mW/m
2
is chosen as that found for our inversion, the
mean glacial temperature is 19 K colder than the
present, which is well in agreement with the results
found for the GRIP reconstruction. Comparison of the
Dye 3 temperature history presented in (7) and that
presented here shows a general good agreement for
the last 7 ky. The history presented in (7) is more
intuitive and less detailed, and the history has not
been corrected for elevation changes. The ice thick-
ness was assumed constant in this reconstruction.
32. L. K. Barlov, J. C. Rogers, M. C. Serreze, R. C. Barry, J.
Geophys. Res. 102, 26333 (1997).
33. R. A. Keen, Occas. Pap. 34 (Institute of Arctic and
Alpine Research, University of Colorado, Boulder,
1980).
34. S. Shubert, W. Higgins, C. K. Park, S. Moorthi, M.
Svarez, An Atlas of ECMWF Analyses (1980–87). Part
II: Second Moment Quantities,NASA Tech. Memoran-
dum 100762 (1990); F. Rex, World Surv. Climatol. 4,
1 (1969).
35. This is a contribution to the Greenland Ice Core
Project (GRIP), a European Science Foundation pro-
gram with eight nations and the European Economic
Commission collaborating to drill through the central
part of the Greenland Ice Sheet. G.D.C. thanks the
USGS Climate History Program and NSF for support.
16 June 1998; accepted 1 September 1998
REPORTS
www.sciencemag.org SCIENCE VOL 282 9 OCTOBER 1998 271
on February 3, 2011www.sciencemag.orgDownloaded from
... m ( Stocker et al. 2013) by the end of the twenty-first century and contributions from the Greenland Ice Sheet (GrIS) are projected to be 8-22 cm by AD 2100 (Bindschadler et al. 2013). Determining the extents and rates of GrIS margin fluctuations during the Holocene, including warmer-than-present times such as the Holocene Thermal Maximum (HTM, 8-5 ka; Briner et al. 2016;Dahl-Jensen et al. 1998;Kaufman et al. 2004), provides a longer-term record of how the ice-sheet responds to climate conditions and may refine predictions of how the ice sheet will respond to future climate changes (e.g., Briner et al. 2010;Carlson et al. 2014;Kelley, Briner, and Zimmerman 2015;Larsen et al. 2014;Levy et al. 2012;Rinterknecht et al. 2009;Winsor et al. 2015;Young et al. 2011). Here, we present sixteen new 10 Be ages of boulders on moraines, boulders perched on bedrock, and bedrock surfaces that con- strain the Holocene extents and thicknesses of the western GrIS margin fluctuations near Kangerlussuaq, Greenland. ...
... Unfortunately, none of these records extend beyond 8 ka, and so it is not possible to evaluate the response of the ice margin to regional climate conditions prior to this time. A chironomid-based summer temperature record from North lake, situated less than 2 km north of Jackobshavn Isfjord and 240 km to the north of our study site, shows relatively warm temperatures from 6 to 3 cal kyr BP and cooling at AD 1880 ( Axford et al. 2013), with an overall Holocene trend that closely resembles the Dye-3 borehole record ( Dahl-Jensen et al. 1998). In com- parison, a qualitative summer temperature record from Lake N3, 55 km south of Jakobshavn Isfjord, which is based on terrestrial leaf wax δ 2 H, shows a cooling trend from 8 to 6 cal kyr BP ( Thomas et al. 2016). ...
Article
Full-text available
The response of the Greenland Ice Sheet to future climate change is relatively unconstrained. Determining the extents and rates of ice-margin fluctuations during the Holocene provides a longer-term perspective on ice-sheet changes and offers an analogue of how the ice-sheet may respond to future changes. Here, we present sixteen new ¹⁰Be ages of boulders on moraines, boulders perched on bedrock, and bedrock surfaces that mark the timing of ice-margin fluctuations during the Holocene in the Kangerlussuaq region of southern west Greenland. We show that the Keglen moraines date to 8.0 ± 0.3 ka (n = 6) and that the average ice-margin retreat rate slowed from about 49 to 13 m yr⁻¹ after about 8.0 ka, likely in response to the ice margin retreating onto land at the head of the fjord Kangerlussuaq at this time. The average retreat rate further slowed to less than 1 m yr⁻¹ between 6.8 ka and 4.2 cal kyr BP, a time when nearby paleoclimate records document warm summers and increased precipitation. Finally, we show that the historical advances of the ice margin occurred during the past 200 years, likely in response to cooler summer temperatures.
... The Holocene climate in Greenland is characterized by an abrupt warming at the glacialinterglacial transition at 11.7 kyr BP, leading to the Holocene climatic optimum at ∼8-5 kyr BP when temperatures were 2-3°C higher than the temperatures in the late 20th century, followed by a generally cooling climate trend until the late 20th century, interrupted with shorter climate fluctuations such as the Little Ice Age (Dahl-Jensen and others, 1998;Kaufman and others, 2004;Gkinis and others, 2014). Records of stable water isotopes, δ 18 O and δD, obtained from deep ice cores are widely used proxies for the past temperature variations across the ice sheet (e.g. ...
... Thus, temperature reconstructions derived directly from the isotopic values of the GRIP record would lack the warming in the early Holocene unless these effects are taken into account. By contrast, temperature histories derived from inversions of borehole temperature (Dahl-Jensen and others, 1998) and recently from isotope-diffusion studies (Gkinis and others, 2014) show a clear signal of the Holocene climatic optimum. The larger temperature signal was also found from the interpretation of the isotopic records of two marginal ice caps, where the proxy record was assumed to be undisturbed by flow and elevation changes (Vinther and others, 2009). ...
Article
Full-text available
The Holocene climatic optimum was a period 8–5 kyr ago when annual mean surface temperatures in Greenland were 2–3°C warmer than present-day values. However, this warming left little imprint on commonly used temperature proxies often used to derive the climate forcing for simulations of the past evolution of the Greenland ice sheet. In this study, we investigate the evolution of the Greenland ice sheet through the Holocene when forced by different proxy-derived temperature histories from ice core records, focusing on the effect of sustained higher surface temperatures during the early Holocene. We find that the ice sheet retreats to a minimum volume of ~0.15–1.2 m sea-level equivalent smaller than present in the early or mid-Holocene when forcing an ice-sheet model with temperature reconstructions that contain a climatic optimum, and that the ice sheet has continued to recover from this minimum up to present day. Reconstructions without a warm climatic optimum in the early Holocene result in smaller ice losses continuing throughout the last 10 kyr. For all the simulated ice-sheet histories, the ice sheet is approaching a steady state at the end of the 20th century.
... Temperature reconstructions from GrIS boreholes and ice cores (Dahl-Jensen and Al, 1998;Buizert, 2014) show that the annual temperature in northern Baffin Bay reached about − 36 • C at the LGM (Jennings et al., 2018). Together with geomorphological evidence for cold-based ice on Baffin Island (Briner et al., 2003), this suggests that the LGM climate was similar to that of present-day Antarctica, and, thus, was cold enough to sustain ice shelves. ...
... Over mid-and highlatitude northern continents, strong reductions in temperatures produced southward displacement and major reductions in forest area (Bigelow et al., 2003), expansion of permafrost limits over northwest Europe (Renssen and Vandenberghe, 2003), fragmentation of temperate forests (Prentice et al., 2000;Williams et al., 2000) and predominance of steppetundra in Western Europe (Peyron et al., 2005). Temperature reconstructions from polar ice cores indicate strong cooling at high latitudes of about 9°C in Antarctica (Stenni et al., 2001) and about 21°C in Greenland (Dahl-Jensen et al., 1998). ...
... This latter record corresponds to the northernmost fossil record found until today, and it has been estimated between 6150 and 5950 years cal BP (Ljungar, 1995). This occurrence thus corresponds with the Atlantic Period (9050-5700 years BP), considered as the climatic optimum of the Holocene, when annual mean temperatures were supposed to be globally 2-2.5°C higher than today (Dahl-Jensen et al., 1998). Besides its mention in some Early Pleistocene (as Z. aff. ...
Article
The Holocene has always been considered a crucial epoch where the major cultural steps of humankind took place. Understanding past Holocene climatic variability, shifts in vegetation, and faunal communities are among crucial challenges in predicting the upcoming changes of the natural environment. In Central Europe, the Kraków-Częstochowa Upland is one of the best study areas for detecting faunal, environmental, and climate shifts. For that reason, we choose to study the small vertebrate and mollusc communities of three archaeological sites with Holocene horizons: Sąspowska Zachodnia Cave, Ciasna Cave, and Małe Rockshelter. We analysed the faunal community shifts, and we compared our results with the major Holocene sites in the area with the correspondence and cluster analyses. The environment and climate were reconstructed through Habitat Weighting, Quantified Ecology, malacological spectra and Bioclimatic methods. Environment and climate values showed a temporary replacement of the Late Glacial tundra and grasslands with boreal forest mixed with open wetlands during the Preboreal and then stable landscapes dominated by temperate forests. Models based on rodent data show that temperature and precipitation generally increased, starting from the Boreal, to reach values near, or sometimes overwhelm, the current climate values of the area during the Atlantic and Subboreal. The climate values obtained from herpetofaunal analyses show very slight changes between the Preboreal and the Atlantic (Ciasna Cave layers 3 and 4), except for the winter temperatures, whereas results obtained for Sąspowska Zachodnia Cave are probably biased by a very low number of recovered remains. Among rodents, a major negative oscillation detected in layers related to the edge of Early Holocene was related to the Preboreal Oscillation or long-lasting effect of the Younger Dryas.
... Shear localization at the base of the ice sheet is enhanced by the softening of ice here due to the geothermal heat flow (Dahl-Jensen et al., 1998). The effect of heating increases at high temperature because the activation energy for ice deformation increases significantly towards its melting point (Barnes et al., 1971;Souchez et al., 1993;Goldsby & Kohlstedt, 2001;Kuiper et al., 2020) and because of the potential presence of water (Gudlaugsson et al., 2016). ...
Article
Full-text available
We present a series of simple shear numerical simulations of dynamic recrystallization of two‐phase nonlinear viscous materials that represent temperate ice. First, we investigate the effect of the presence of water on the resulting microstructures and, second, how water influences on P wave (V p ) and fast S wave (V s ) velocities. Regardless the water percentage, all simulations evolve from a random fabric to a vertical single maximum. For a purely solid aggregate, the highest V p quickly aligns with the maximum c‐axis orientation. At the same time, the maximum c‐axis development reduces V s in this orientation. When water is present, the developed maximum c‐axis orientation is less intense, which results in lower V p and V s . At high percentage of water, V p does not align with the maximum c‐axis orientation. If the bulk modulus of ice is assumed for the water phase (i.e., implying that water is at high pressure), we find a remarkable decrease of V s while V p remains close to the value for purely solid ice. These results suggest that the decrease in V s observed at the base of the ice sheets could be explained by the presence of water at elevated pressure, which would reside in isolated pockets at grain triple junctions. Under these conditions water would not favor sliding between ice grains. However, if we consider that deformation dominates over recrystallization, water pockets get continuously stretched, allowing water films to be located at grain boundaries. This configuration would modify and even overprint the maximum c‐axis‐dependent orientation and the magnitude of seismic anisotropy.
... The basal temperature at the GISP2 and GRIP ice core sites are reportedly cold at ∼−9°C (Dahl- Jensen and others, 1998). Both radar datasets detect water a few tens of kilometers to the east of these ice cores. ...
Article
Full-text available
The spatial distribution of basal water critically impacts the evolution of ice sheets. Current estimates of basal water distribution beneath the Greenland Ice Sheet (GrIS) contain large uncertainties due to poorly constrained boundary conditions, primarily from geothermal heat flux (GHF). The existing GHF models often contradict each other and implementing them in numerical ice-sheet models cannot reproduce the measured temperatures at ice core locations. Here we utilize two datasets of radar-detected basal water in Greenland to constrain the GHF at regions with a thawed bed. Using the three-dimensional ice-sheet model SICOPOLIS, we iteratively adjust the GHF to find the minimum GHF required to reach the bed to the pressure melting point, GHF pmp , at locations of radar-detected basal water. We identify parts of the central-east, south and northwest Greenland with significantly high GHF pmp . Conversely, we find that the majority of low-elevation regions of west Greenland and parts of northeast have very low GHF pmp . We compare the estimated constraints with the available GHF models for Greenland and show that GHF models often do not honor the estimated constraints. Our results highlight the need for community effort to reconcile the discrepancies between radar data, GHF models, and ice core information.
Article
As we struggle to cope with the ongoing buildup of CO2 produced by burning fossil fuels, can we acquire guidance from the geologic record? Although our ability to reconstruct past atmospheric CO2 content reliably is currently confined to the last 800 thousand years, we do have compelling evidence that this greenhouse gas played a key role throughout the Earth's history. It certainly compensated for the young Sun's lower luminosity. There is no question that it bailed us out of two snowball episodes or that it led to a brief 5°C warming at the onset of the Eocene. Less certain is that diminishing atmospheric CO2 content was responsible for the global cooling that began 50 million years ago when the Indian subcontinent collided with Asia. Finally, it colluded with changing seasonality, ocean circulation re-organisation and iron fertilisation to generate the 100 thousand year glacial cycles that dominated the last half-million years. © 2018 European Association of Geochemistry. All rights reserved.
Article
Full-text available
We describe a high-precision (0.1–1.0 mK) borehole-temperature (BT) logging system developed at the United States Geological Survey (USGS) for use in remote polar regions. We discuss calibration, operational and data-processing procedures, and present an analysis of the measurement errors. The system is modular to facilitate calibration procedures and field repairs. By interchanging logging cables and temperature sensors, measurements can be made in either shallow air-filled boreholes or liquid-filled holes up to 7 km deep. Data can be acquired in either incremental or continuous-logging modes. The precision of data collected by the new logging system is high enough to detect and quantify various thermal effects at the milli-Kelvin level. To illustrate this capability, we present sample data from the 3 km deep borehole at GISP2, Greenland, and from a 130 m deep air-filled hole at Taylor Dome, Antarctica. The precision of the processed GISP2 continuous temperature logs is 0.25–0.34 mK, while the accuracy is estimated to be 4.5 mK. The effects of fluid convection and the dissipation of the thermal disturbance caused by drilling the borehole are clearly visible in the data. The precision of the incremental Taylor Dome measurements varies from 0.11 to 0.32 mK. depending on the wind strength during the experiments. With this precision, we found that temperature fluctuations and multi-hour trends in the BT measurements correlate well with atmospheric-pressure changes.
Article
Full-text available
We present a combined heat- and ice-flow model, constrained by measurements of temperature in the Greenland Ice Sheet Project 2 (GISP2) borehole and by the GISP2 51so record and depth-age scale, which determines a history of temperature, accumulation rate, and ice sheet elevation for the past 50,000 years in central Greenland. Important results are: that the temperature increase from average glacial to Holocene conditions was large, approximately 15 øC, with a 20 øC warming from late glacial to Holocene; that the average accumulation rate during the last glacial maximum (between 15 and 30 kyr B. P.) was 5.5 to 7 cm yr -1, approximately 25% of the modern accumulation rate; that long-term (500-1000 years) averaged accumulation rate and temperature have been inversely correlated during the most recent 7 millennia of the Holocene; and that the Greenland Ice Sheet probably thickened during the aleglacial transition. The inverse correlation of accumulation rate and temperature in the mid and late Holocene suggests that the Greenland Ice Sheet is more prone to volume reduction in a warmed climate than previously thought and demonstrates that accumulation rate is not a reliable proxy for temperature. The elevation history of the ice sheet is poorly constrained by the model, and independent evidence is needed. We also present a simple estimate of the response time for thinning of the interior region of an ice sheet due to retreat of its margins. This was approximately 1900 years for central Greenland during deglaciation.
Article
Meltwater influxes may partly explain the low oxygen-isotope values measured in the Dye 3 and Camp Century ire cores. This has led to speculation that Greenland may not have cooled during the Younger Dryas and underlined the need for independent checks of the oxygen-isotope record. Using optimal control methods and heat-flow modeling, the author makes a valiant but ultimately futile attempt to distinguish the Younger Dryas event in the ice-sheet temperatures measured at Dye 3, south Greenland. The author discusses the prospects for attempting the same in the new Summit boreholes in central Greenland: how that will require more accurate temperature measurements, a coupled thermo-mechanical model and a refined uncertainty analysis. He concludes by discussing how borehole-temperature analysis may improve the climate histories determined from ire cores.
Article
In 1983 three directional surveys were made in the bore hole from which a deep ice core was obtained in the summers of 1979–81. The inclination and azimuth of the bore hole were measured on three surveys, temperature was included on two surveys, fluid pressure and hole diameter on one of the surveys. Fluid-pressure measurements show that the ice-overburden pressure was undercompensated in the upper few hundred meters and overcompensated at the bottom of the hole. Diameter measurements show closure in the upper portion and expansion near the bottom beginning at the transition from the Holocene to Wisconsin ice at 1784 m. The hole expansion and increase in inclination correlate with dust and silt content in the Wisconsin ice. Changes in azimuth are due to flow of the ice and are consistent with the direction of flow at the surface. Temperature measurements show that the hole is at or near equilibrium. The gradient of 0.012 K/m below 1400 m is less than the 0.018 K/m at Camp Century. There is a slight reduction in gradient near the bottom from internal friction in the silty ice.
Article
Heat flow and surface radioactivity were determined at two sites in the Precambrian shield of south Greenland. At Ivigtut heat flow q from two holes in Ketilidian (˜1700-1500 m.y.) gneisses averaged 1.0 ± 0.1 HFU (μcal cm-2 sec-1), and the radiogenic heat production A0 averaged 5.5 ± 1.5 HGU (10-13 cal cm-3 sec-1). These data are consistent with other pairs of q and A0 values from the Canadian and Australian shields. In four holes near the northwestern edge of the Ilímaussaq alkali intrusion of Gardar (˜1000 m.y.) age, however, q averages only 0.9 ± 0.1 HFU, whereas A0 averages >20 HGU. This observation implies that, unless heat flow from the mantle is unexpectedly low in this part of Greenland, the highly radioactive part of the intrusion is quite thin (˜1 km) and the underlying crustal rocks contain very little radioactivity.
Article
Static and dynamic equilibrium properties of butane octane, and dodecane films adsorbed on α-Al2O3(0001) at a variety of coverages and temperatures, and the subsequent penetration of such films by 30 molecule water clusters are examined using classical molecular dynamics. Model potential functions are constructed from existing alkane united atom and “simple point charge” model water parameters, experimental alkane desorption energies and other available theoretical information. The adsorbed films exhibit a distinct layering parallel to the surface, and a pronounced densification, reduction in gauche defects and orientational ordering within the innermost layer. Strong surface corrugation allows molecules to rotate relatively freely about their long axes at intermediate temperatures and assists them in orienting their zig-zag planes perpendicular to the surface at lower temperatures. Only butane molecules show any tendency to tilt their long axes out of the first layer toward the second. (H2O)30 clusters are attracted toward the alumina surface and easily penetrate most of the adsorbed alkane films, either by displacing alkane molecules to more distant layers or causing them to pack more closely within existing layers. The molecules in the clusters tend to remain connected during penetration. Kinetic barriers to penetration become increasingly significant for higher alkane coverages, lower temperatures, and longer chains.
Article
The Car–Parrinello method has been applied to study the adsorption of water on solid magnesium oxide with surface defects. A step consisting of an (100) and an (010) surface on an (011) base plane allows us to model the experimentally observed microfaceting. In and on this step dissociation of water into a hydroxyl group and a H‐atom took place following a complicated pathway only accessible by the simulation of thermal motion. Under comparable conditions physisorption only was observed on a regular (001) plane. This solves an experimental controversy and it is in agreement with the observation, that disordered surfaces are more active in initiating the dissociation of the water molecules. Our work allows us to identify an important active center. We can also account for the experimentally observed broadening and shifting to the red of the stretching mode of hydrogen bonded hydroxyl groups, and we provide a detailed explanation of the origin of this effect. This allows us to verify earlier theories of hydrogen bonding such as that of the adiabatic separation of the proton dynamics. © 1995 American Institute of Physics.
Article
The anthraquinone dye molecule quinizarin is known to allow for persistent spectral hole burning up to liquid nitrogen temperatures after chemisorption on alumina surfaces. The mechanism underlying these improved hole‐burning properties is not known, though is has been speculated that it might be related to intrinsic surface effects on the electronic structure of the dye. We approach this problem theoretically using gradient corrected density functional theory. The chemisorbed compound system is modelled by a periodically replicated nine layer slab which represents the (0001) surface of α‐Al2O3. The chemisorption geometry obtained by geometry optimization and confirmed by Car–Parrinello molecular dynamics runs at room temperature is shown to be a perpendicular arrangement of quinizarin on the surface, where a chelate‐like bond is formed with one exposed surface aluminum atom. In order to get information about the electronic structure, the frontier orbitals that are relevant for the description of the electronic excitation to the first excited state are evaluated for the isolated molecule, the chemisorbed molecule, and a quinizarin‐aluminum‐water complex. The strong red shift of the absorption frequency found in experiment upon chemisorption is reproduced. However, the results show that the shape of the frontier orbitals and hence the properties of the electronic excitation remain essentially unchanged by chemisorption. Thus, the differences in the behavior of the isolated and the chemisorbed dye observed in persistent spectral hole‐burning experiments cannot be explained by genuine surface induced effects on the molecular electronic structure. © 1996 American Institute of Physics.
Article
Hydroxylated (0001) faces of α-Al2O3 bulk crystals have been examined by high-resolution electron-energy loss spectroscopy (HREELS). Surface charging being eliminated by means of a defocused electron beam, it is shown that the electron fluence can be low enough to avoid significant damages. A resonance scattering from chemisorbed OH groups is evidenced, at impact energies lower than 12 eV, by strong enhancements of the cross sections for excitation of both the fundamental (3720 cm−1) and the first overtone (7270 cm−1) of the stretching mode and by the angular distribution of the inelastically scattered electrons.