ArticlePDF Available

Temperature-driven global sea-level variability in the Common Era

Authors:

Abstract and Figures

Significance We present the first, to our knowledge, estimate of global sea-level (GSL) change over the last ∼3,000 years that is based upon statistical synthesis of a global database of regional sea-level reconstructions. GSL varied by ∼±8 cm over the pre-Industrial Common Era, with a notable decline over 1000–1400 CE coinciding with ∼0.2 °C of global cooling. The 20th century rise was extremely likely faster than during any of the 27 previous centuries. Semiempirical modeling indicates that, without global warming, GSL in the 20th century very likely would have risen by between −3 cm and +7 cm, rather than the ∼14 cm observed. Semiempirical 21st century projections largely reconcile differences between Intergovernmental Panel on Climate Change projections and semiempirical models.
Content may be subject to copyright.
Temperature-driven global sea-level variability in the
Common Era
Robert E. Kopp
a,b,c,1
, Andrew C. Kemp
d
, Klaus Bittermann
e
, Benjamin P. Horton
b,f,g,h
, Jeffrey P. Donnelly
i
,
W. Roland Gehrels
j
, Carling C. Hay
a,b,k
, Jerry X. Mitrovica
k
, Eric D. Morrow
a,b
, and Stefan Rahmstorf
e
a
Department of Earth & Planetary Sciences, Rutgers University, Piscataway, NJ 08854;
b
Institute of Earth, Ocean & Atmospheric Sciences, Rutgers University,
New Brunswick, NJ 08901;
c
Rutgers Energy Institute, Rutgers University, New Brunswick, NJ 08901;
d
Department of Earth & Ocean Sciences, Tufts University,
Medford, MA 02115;
e
Earth System Analysis, Potsdam Institute for Climate Impact Research, 14473 Potsdam, Germany;
f
Sea-Level Research, Department
of Marine & Coastal Sciences, Rutgers University, New Brunswick, NJ 08901;
g
Earth Observatory of Singapore, Nanyang Technological University,
Singapore 639798;
h
Asian School of the Environment, Nanyang Technological University, Singapore 639798;
i
Department of Geology and Geophysics,
Woods Hole Oceanographic Institution, Woods Hole, MA 02543;
j
Environment Department, University of York, York YO10 5NG, United Kingdom; and
k
Department of Earth & Planetary Sciences, Harvard University, Cambridge, MA 02138
Edited by Anny Cazenave, Centre National dEtudes Spatiales, Toulouse, France, and approved January 4, 2016 (received for review August 27, 2015)
We assess the relationship between temperature and global sea-
level (GSL) variability over the Common Era through a statistical
metaanalysis of proxy relative sea-level reconstructions and tide-
gauge data. GSL rose at 0.1 ±0.1 mm/y (2σ) over 0700 CE. A GSL
fall of 0.2 ±0.2 mm/y over 10001400 CE is associated with 0.2 °C
global mean cooling. A significant GSL acceleration began in the
19th century and yielded a 20th century rise that is extremely
likely (probability P0.95) faster than during any of the previous
27 centuries. A semiempirical model calibrated against the GSL
reconstruction indicates that, in the absence of anthropogenic cli-
mate change, it is extremely likely (P=0.95) that 20th century
GSL would have risen by less than 51% of the observed 13.8 ±1.5 cm.
The new semiempirical model largely reconciles previous differ-
ences between semiempirical 21st century GSL projections and
the process model-based projections summarized in the Inter-
governmental Panel on Climate Changes Fifth Assessment
Report.
sea level
|
Common Era
|
late Holocene
|
climate
|
ocean
Estimates of global mean temperature variability over the
Common Era are based on global, statistical metaanalyses of
temperature proxies (e.g., refs. 13). In contrast, reconstructions
of global sea-level (GSL) variability have relied upon model
hindcasts (e.g., ref. 4), regional relative sea-level (RSL) recon-
structions adjusted for glacial isostatic adjustment (GIA) (e.g.,
refs. 58), or iterative tuning of global GIA models (e.g., ref. 9).
Based primarily on one regional reconstruction (8), the In-
tergovernmental Panel on Climate Change (IPCC)s Fifth As-
sessment Report (AR5) (10) concluded with medium confidence
that GSL fluctuations over the last 5 millennia were 25 cm.
However, AR5 was unable to determine whether specific fluc-
tuations seen in some regional records (e.g., ref. 5) were global in
extent. Similarly, based upon a tuned global GIA model, ref. 9
found no evidence of GSL oscillations exceeding 1520 cm
between 2250 and 1800 CE and no evidence of GSL trends
associated with climatic fluctuations.
The increasing availability and geographical coverage of con-
tinuous, high-resolution Common Era RSL reconstructions
provides a new opportunity to formally estimate GSL change over
the last 3,000 years. To do so, we compiled a global database of
RSL reconstructions from 24 localities (Dataset S1, a and Fig.
S1A), many with decimeter-scale vertical resolution and sub-
centennial temporal resolution. We augment these geological re-
cords with 66 tide-gauge records, the oldest of which (11) begins in
1700 CE (Dataset S1,bandFig. S1B), as well as a recent tide-
gaugebased estimate of global mean sea-level change since 1880
CE (12).
To analyze this database, we construct a spatiotemporal em-
pirical hierarchical model (13, 14) that distinguishes between sea-
level changes that are common across the database and those
that are confined to smaller regions. The RSL field fðx,tÞis
represented as the sum of three components, each with a Gaussian
process (GP) prior (15),
fðx,tÞ=gðtÞ+lðxÞðtt0Þ+mðx,tÞ.[1]
Here, xrepresents spatial location, trepresents time, and t0is a
reference time point (2000 CE). The three components are (i)
GSL gðtÞ, which is common across all sites and primarily repre-
sents contributions from thermal expansion and changing land
ice volume; (ii) a regionally varying, temporally linear field
lðxÞðtt0Þ, which represents slowly changing processes such as
GIA, tectonics, and natural sediment compaction; and (iii)a
regionally varying, temporally nonlinear field mðx,tÞ, which pri-
marily represents factors such as ocean/atmosphere dynamics
(16) and static equilibrium fingerprinteffects of landice mass
balance changes (17, 18). The regional nonlinear field also
incorporates small changes in rates of GIA, tectonics, and
compaction that occur over the Common Era. The incorpora-
tion of the regionally correlated terms lðxÞðtt0Þand mðx,tÞ
ensures that records from regions with a high density of
Significance
We present the first, to our knowledge, estimate of global
sea-level (GSL) change over the last 3,000 years that is
based upon statistical synthesis of a global database of re-
gional sea-level reconstructions. GSL varied by ±8cmover
the pre-Industrial Common Era, with a notable decline over
10001400 CE coinciding with 0.2 °C of global cooling. The
20th century rise was extremely likely faster than during any
of the 27 previous centuries. Semiempirical modeling indi-
cates that, without global warming, GSL in the 20th century
very likely would have risen by between 3cmand+7cm,
rather than the 14 cm observed. Semiempirical 21st century
projections largely reconcile differences between Intergov-
ernmental Panel on Climate Change projections and semi-
empirical models.
Author contributions: R.E.K. designed research; R.E.K., A.C.K., K.B., B.P.H., J.P.D., and
W.R.G. performed research; R.E.K., K.B., C.C.H., J.X.M., E.D.M., and S.R. contributed new
analytic tools; R.E.K. and K.B. analyzed data; R.E.K., A.C.K., K.B., B.P.H., J.P.D., W.R.G., C.C.H.,
J.X.M., E.D.M., and S.R. wrote the paper; A.C.K., B.P.H., and W.R.G. compiled the
database of proxy reconstructions; C.C.H., J.X.M., and E.D.M. contributed to the design of
the statistical model; and K.B. and S.R. developed and implemented the semiempirical
projections.
The authors declare no conflict of interest.
This article is a PNAS Direct Submission.
Freely available online through the PNAS open access option.
1
To whom correspondence should be addressed. Email: robert.kopp@rutgers.edu.
This article contains supporting information online at www.pnas.org/lookup/suppl/doi:10.
1073/pnas.1517056113/-/DCSupplemental.
www.pnas.org/cgi/doi/10.1073/pnas.1517056113 PNAS Early Edition
|
1of8
SUSTAINABILITY
SCIENCE
PNAS PLUS
observations are not unduly weighted in estimating the common
GSL signal gðtÞ.
Because a constant-rate trend in gðtÞcould also be interpreted as
a regional linear trend that is present at all reconstruction sites but
is not truly global, we condition the model on the assumption that
mean GSL over 100100 CE is equal to mean GSL over 1600
1800 CE and focus on submillennial variations (Fig. 1A). We chose
the first window to encompass the beginning of the Common Era
and the last window to cover the last 2 centuries before the de-
velopment of a tide-gauge network outside of northern Europe.
The priors for each component are characterized by hyper-
parameters that comprise amplitudes (for all three components),
timescales of variability [for gðtÞand mðx,tÞ], and spatial scales
of variability [for lðxÞand mðx,tÞ](Dataset S1, c). We consider
five priors with different hyperparameters (see Supporting In-
formation). The presented rates are taken from prior ML
2,1
,
which is optimized under the assumption that the a priori
timescales of variability in global and regional sea-level change
are the same. Results from the four alternative priors are pre-
sented in Supporting Information. Quoted probabilities are con-
servatively taken as minima across all five priors. Illustrative fits
at specific sites are shown in Fig. S2.
Results and Discussion
Common Era Reconstruction. Pre-20th-century Common Era GSL
variability was very likely (probability P=0.90) between ±7cm
and ±11 cm in amplitude (Fig. 1Aand Dataset S1, e). GSL rose
from 0 CE to 700 CE (P0.98) at a rate of 0.1 ±0.1 mm/y (2σ),
was nearly stable from 700 CE to 1000 CE, then fell from 1000
CE and 1400 CE (P0.98) at a rate of 0.2 ±0.2 mm/y (Fig. 1A).
GSL likely rose from 1400 CE to 1600 CE (P0.75) at 0.3 ±
0.4 mm/y and fell from 1600 CE to 1800 CE (P0.86) at
0.3 ±0.3 mm/y.
Historic GSL rise began in the 19th century, and it is very
likely (P0.93) that GSL has risen over every 40-y interval
sin ce 1 860 CE. The average rate of GSL rise was 0.4 ±0.5 mm/y
from 1860 CE to 1900 CE and 1.4 ±0.2 mm/y over the 20th
century. It is extremely likely (P0.95) that 20th century GSL
rise was faster than during any preceding century since at least
800 CE.
The spatial coverage of the combined proxy and long-term
tide-gauge dataset is incomplete. The available data are suffi-
cient to reduce the posterior variance in the mean 01700 CE
rate by >10% relative to the prior variance along coastlines in
much of the North Atlantic and the Gulf of Mexico, and parts of
the Mediterranean, the South Atlantic, the South Pacific, and
Australasia (Fig. 2A). High-resolution proxy records are notably
lacking from Asia, most of South America, and most of Africa.
Nevertheless, despite the incomplete coverage and regional
variability, sensitivity analyses of different data subsets indicate
that key features of the GSL curvea rise over 0700 CE, a fall
over 10001400CE,andarisebeginninginthelate19th
centuryare not dependent on records from any one region
(Dataset S1, f). By contrast, the rise over 14001600 CE and fall
over 16001800 CE are not robust to the removal of data from
the western North Atlantic.
On millennial and longer timescales, regional RSL change can
differ significantly from GSL change as a result of GIA, tec-
tonics, and sediment compaction (Fig. 2). For example, over 0
1700 CE, RSL rose at 1.5 ±0.1 mm/y in New Jersey, on the
collapsing forebulge of the former Laurentide Ice Sheet, and fell
at 0.1 ±0.1 mm/y on Christmas Island, in the far field of all late
Pleistocene ice cover (Dataset S1, g). Detrended RSL (after re-
moval of the average 01700 CE rate) reveals notable patterns of
temporal variability, especially in the western North Atlantic,
where the highest-resolution reconstructions exist. Rates of RSL
change in New Jersey and North Carolina vary from the long-
term mean in opposite directions over 0700 CE and 10001400
CE (Fig. 2 and Dataset S1, g). Over 0700 CE, a period over
which GSL rose at 0.1 ±0.1 mm/y, detrended RSL rose in New
Jersey (P0.91) while it fell in North Carolina (P0.88). Con-
versely, over 10001400 CE, while GSL was falling, detrended
RSL fell in New Jersey (P>0.90) while it rose in North Car-
olina (P0.99). This pattern is consistent with changes in the
Gulf Stream (16) or in mean nearshore wind stress (19). If driven
by the Gulf Stream, it suggests a weakening or polar migration of
the Gulf Stream over 0700 CE, with a strengthening or equatorial
migration occurring over 10001400 CE.
Global SeaLevel (cm)
C
Year (CE)
(change in scale)
E
Marcott et al. (2013)Global sea level (this study) Mann et al. (2009)
D
RCP 8.5
RCP 4.5
RCP 2.6
Maximum range across all RCPs and calibrations
F
RCP 2.6
RCP 4.5
RCP 8.5
1850 1900 1950 2000 2050 2100
-20
0
20
40
60
80
100
120
B
1
2
H
A
Global Temperature Anomaly (oC) Global Sea Level (cm)GlobalSea Level (c m)
2000
-400 - 200 0 200 400 600 800 1000 1200 1400 1600 1800
-10
-5
0
5
10
15
20
-400 -200 0 200 400 600 800 1000 1200 1400 1600 1800 2000
-1.0
-0.8
-0.6
-0.4
-0.2
0.0
0.2
0.4
0.6
-400 -200 0 200 400 600 800 1000 1200 1400 1600 1800 2000
-10
-5
0
5
10
15
20
Fig. 1. (A) Global sea level (GSL) under prior ML
2,1
. Note that the model is
insensitive to small linear trends in GSL over the Common Era, so the relative
heights of the 3001000 CE and 20th century peaks are not comparable. (B)
The 90% credible intervals for semiempirical hindcasts of 20th century sea-
level change under historical temperatures (H) and counterfactual scenarios
1 and 2, using both temperature calibrations. (C) Reconstructions of global
mean temperature anomalies relative to the 18502000 CE mean (1, 2). (D)
Semiempirical fits to the GSL curve using the two alternative temperature
reconstructions. (E)AsinB, including 21st century projections for RCPs 2.6,
4.5, and 8.5. Red lines show the fifth percentile of RCP 2.6 and 95th per-
centile of RCP 8.5. (F) The 90% credible intervals for 2100 by RCP. In A,B, and
D, values are with respect to 1900 CE baseline; in Eand F, values are with
respect to 2000 CE baseline. Heavy shading, 67% credible interval; light
shading, 90% credible interval.
2of8
|
www.pnas.org/cgi/doi/10.1073/pnas.1517056113 Kopp et al.
Our estimate differs markedly from previous reconstructions of
Common Era GSL variability (5, 6, 9, 20) (Fig. S3F). For example,
the ref. 20 hindcast predicts GSL swings with 4×larger amplitude,
and it includes a rise from 650 CE to 1200 CE (a period of GSL
stability and fall in the data-based estimate) and a fall from 1400
CE to 1700 CE (a period of approximate GSL stability in the
data-based estimate). The curve derived from the detrended North
Carolina RSL reconstruction (5) indicates an amplitude of change
closer to our GSL reconstruction but differs in phasing from it, with
a relatively high sea level during 12001500 CE likely reflecting
the regional processes mentioned above. The globally tuned GIA
model of ref. 9, which includes 31 data points from the last mil-
lennium (compared with 790 proxy data points in our analysis),
found no systematic GSL changes over the Common Era.
Twentieth Century GSL Rise. Semiempirical models of GSL change,
based upon statistical relationships between GSL and global mean
temperature or radiative forcing, provide an alternative to process
models for estimating future GSL rise (e.g., refs. 2023) and
generating hypotheses about past changes (e.g., refs. 4, 20, and
24). The underlying physical assumption is that GSL is expected to
rise in response to climatic warming and reach higher levels during
extended warm periods, and conversely during cooling and ex-
tended cool periods. Ref. 5 generated the first semiempirical GSL
model calibrated to Common Era proxy data, but relied upon sea-
level data from a single region rather than a global synthesis.
Our new GSL curve shows that multicentury GSL variability
over the Common Era shares broad commonalities with global
mean temperature variability, consistent with the assumed link
that underlies semiempirical models. For example, the 9 ±8cm
GSL fall over 10001400 CE coincides with a 0.2 °C decrease in
global mean temperature, and the 9 ±3 cm GSL rise over 1860
1960 CE coincides with 0.2 °C warming (2). Motivated by these
commonalities, using our GSL reconstruction and two global
mean temperature reconstructions (1, 2), we construct a semi-
empirical GSL model that is able to reproduce the main features
of GSL evolution (Fig. 1 Band C).
To assess the anthropogenic contribution to GSL rise, we
consider two hypothetical global mean temperature scenarios
without anthropogenic warming. In scenario 1, the gradual
temperature decline from 500 CE to 1800 CE is taken as rep-
resentative of Earths long-term, late Holocene cooling (2), and,
in 1900 CE, temperature returns to a linear trend fit to 5001800
CE. In scenario 2, we assume that 20th century temperature
stabilizes at its 5001800 CE mean. The difference between GSL
change predicted under these counterfactuals and that predicted
0-1700 CE
0.0
-1.0
1.0
2.0
3.0
0-700 CE
0.0
0.1
0.2
GSL: 0.1 ± 0.1 mm/y
-0.3
-0.2
-0.1
0.0
0.1
GSL: -0.1 ± 0.1 mm/y
-0.1
-0.2
0.0
0.1
0.2
GSL: 0.0 ± 0.2 mm/y
-0.6
-0.4
-0.2
0.0
0.2
0.4
0.6
GSL: 0.0 ± 0.4 mm/y
0.5
0.0
1.0
1.5
2.0
2.5
GSL: 1.4 ± 0.2 mm/y
AB
CD
EF
700-1400 CE 1400-1800 CE
1900-2000 CE
1800-1900 CE
Fig. 2. (A) Mean estimated rate of change (millimeters per year) over 01700 CE under prior ML
2,1
. In shaded areas, conditioning on the observations reduces
the variance by at least 10% relative to the prior. (BF) Mean estimated rates of change (mm/y) from (B)0700 CE, (C) 7001400 CE, (D) 14001800 CE, (E)
18001900 CE, and (F) 19002000 CE, after removing the 01700 CE trend. Areas where a rise and a fall are about equally likely (P=0.330.67) are cross-
hatched. The color scales are centered around the noted rate of GSL change.
Kopp et al. PNAS Early Edition
|
3of8
SUSTAINABILITY
SCIENCE
PNAS PLUS
under observed temperatures represents two alternative interpre-
tations of the anthropogenic contribution to GSL rise (Table 1,
Fig. 1A, and Fig. S4). Both scenarios show a dominant human
influence on 20th century GSL rise.
The hindcast 20
th
century GSL rise, driven by observed tem-
peratures, is 13 cm, with a 90% credible interval of 7.717.5 cm.
This is consistent with the observed GSL rise of 13.8±1.5 cm,
which is due primarily to contributions from thermal expansion
and glacier mass loss (25). Of the hindcast 20th century GSL rise,
it is very likely (P=0.90) that 27% to +41% of the total
(scenario 1) or 10% to +51% of the total (scenario 2) would
have occurred in the absence of anthropogenic warming. Under
all calibrations and scenarios, it is likely (P0.88) that observed
20th century GSL rise exceeded the nonanthropogenic counter-
factuals by 1940 CE and extremely likely (P0.95) that it had
done so by 1950 CE (Dataset S1, h). The GSL rise in the alter-
native scenarios is related to the observation that global mean
temperature in the early 19th century was below the 5001800 CE
trend (and thus below the 20th century in scenario 1) and, for most
of the 19th century, was below the 5001800 CE mean (and thus
below the 20th century in scenario 2) (Fig. S4).
The estimates of the nonanthropogenic contribution to
20th century GSL rise are similar to ref. 4s semiempirical
estimate of 17 cm. They are also comparable to the detrended
fluctuation analysis estimates of refs. 26 and 27, which found it
extremely likely that <40% of observed GSL rise could be
explained by natural variability. These previous estimates, how-
ever, could have been biased low by the short length of the record
used. The 3,000-y record underlying our estimates provides
greater confidence.
Projected 21st Century GSL Rise. The semiempirical model can be
combined with temperature projections for different Represen-
tative Concentration Pathways (RCPs) to project future GSL
change (Table 2, Fig. 1D, and Dataset S1, i). RCPs 8.5, 4.5, and
2.6 correspond to high-end business-as-usualgreenhouse gas
emissions, moderate emissions abatement, and extremely strong
emissions abatement, respectively. They give rise to very likely
(P=0.90) GSL rise projections for 2100 CE (relative to 2000 CE)
of 52131 cm, 3385 cm, and 2461 cm, respectively. Comparison
of the RCPs indicates that a reduction in 21st century sea-level
rise of 30 to 70 cm could be achieved by strong mitigation
efforts (RCP 2.6), even though sea level is a particularly slow-
respondingcomponent of the climate system.
Since ref. 21 inaugurated the recent generation of semi-
empirical models with its critique of the process model-based GSL
projections of the IPCCs Fourth Assessment Report (AR4) (28),
semiempirical projections have generally exceeded those based
upon process models. While AR5s projections (29) were signifi-
cantly higher than those of AR4, semiempirical projections (e.g.,
ref. 23) have continued to be higher than those favored by the
IPCC. However, our new semiempirical projections are lower
than past results, and they overlap considerably with those of
AR5 (29) and of ref. 30, which used a bottom-up probabilistic
estimate of the different factors contributing to sea-level
change. They also agree reasonably well with the expert sur-
vey of ref. 31 (Table 2). Our analysis thus reconciles the
remaining differences between semiempirical and process-
based models of 21st century sea-level rise and strengthens
confidence in both sets of projections. However, both semi-
empirical and process model-based projections may un-
derestimate GSL rise if new processes not active in the
calibration period and not well represented in process models
[e.g., marine ice sheet instability in Antarctica (32)] become
major factors in the 21st century.
Conclusions
We present, to our knowledge, the first Common Era GSL re-
construction that is based upon the statistical integration of a
global database of RSL reconstructions. Estimated GSL variability
Table 1. Hindcasts of 20th century GSL rise (centimeters)
Scenario
Summary
Calibrated to individual temperature reconstructions
Mann et al. (1) Marcott et al. (2)
50th percentile 5th95th percentile 50th percentile 5th95th percentile 50th percentile 5th95th percentile
Observed 13.8 12.615.0
Historical 13.0 7.717.5 14.3 11.517.5 11.6 7.716.3
Scenario 1 0.6 3.54.1 0.9 1.33.3 0.3 3.54.1
Scenario 2 4.0 0.97.5 5.5 3.37.5 2.4 0.95.9
Percent of historical
Scenario 1 5 2741 7 923 3 2741
Scenario 2 30 1051 39 2451 21 1042
All values are with respect to year 1900 CE baseline. Summary results show means of medians, minima of lower bounds, and maxima of upper bounds
taken across both temperature calibrations.
Table 2. Projections of 21st century GSL rise (centimeters)
Method
This study
semiempirical
AR5 (29)
assessment
Schaeffer et al. (23)
semiempirical
Kopp et al. (30)
bottom-up
Horton et al. (31)
survey
50th
percentile
17th83rd
percentile
5th95th
percentile
50th
percentile
17th83rd
percentile
50th
percentile
5th95th
percentile
50th
percentile
17th83rd
percentile
5th95th
percentile
17th83rd
percentile
5th95th
percentile
RCP 2.6 38 2851 2461 43 2860 75 5296 50 3765 2982 4060 2570
RCP 4.5 51 3969 3385 52 3570 90 64121 59 4577 3693 n.a. n.a.
RCP 8.5 76 59105 52131 73 5397 n.c. n.c. 79 62100 55121 70120 50150
All values are with respect to year 2000 CE baseline except AR5, which is with respect to the 19852005 CE average. Results from this study show mean of
medians, minima of lower bounds, and maxima of upper bounds. n.a., not asked; n.c., not calculated.
4of8
|
www.pnas.org/cgi/doi/10.1073/pnas.1517056113 Kopp et al.
over the pre-20th century Common Era was very likely between
±7 cm and ±11 cm, which is more tightly bound than the
25 cm assessed by AR5 (10) and smaller than the variability
estimated by a previous semiempirical hindcast (4). The most
robust pre-Industrial signals are a GSL increase of 0.1 ±0.1
mm/y from 0 CE to 700 CE and a GSL fall of 0.2 ±0.2 mm/y
from 1000 CE to 1400 CE. The latter decline coincides with a
decline in global mean temperature of 0.2 °C, motivating the
construction of a semiempirical model that relates the rate of
GSL change to global mean temperature. Counterfactual hind-
casts with this model indicate that it is extremely likely (P=0.95)
that less than about half of the observed 20th century GSL rise
would have occurred in the absence of global warming, and that
it is very likely (P=0.90) that, without global warming, 20th
century GSL rise would have been between 3 cm and +7 cm,
rather than the observed 14 cm. Forward projections indicate a
very likely 21st century GSL rise of 52131 cm under RCP 8.5
and 2461 cm under RCP 2.6, values that provide greater con-
sistency with process model-based projections preferred by AR5
than previous semiempirical projections.
Materials and Methods
Sea-Level Records. The database of RSL reconstructions (Dataset S2) was com-
piled from published literature, either directly from the original publications or
by contacting the corresponding author (5, 7, 8, 3389). The database is not a
complete compilation of all sea-level index points from the last 3,000 years.
Instead, we include only those reconstructions that we qualitatively assessed as
having sufficient vertical and temporal resolution and density of data points to
allow identification of nonlinear variations, should they exist. This assessment
was primarily based on the number of independent age estimates in each re-
cord. Where necessary and possible, we also included lower-resolution recon-
structions to ensure that long-term linear trends were accurately captured if the
detailed reconstruction was of limited duration. For example, the detailed re-
construction from the Isle of Wight (69) spans only the last 300 y, and we
therefore included a nearby record that described regional RSL trends in
southwest England over the last 2,000 y (51).
Each database entry includes reconstructed RSL, RSL error, age, and age
error. For regional reconstructions produced from multiple sites (e.g., ref. 5),
we treated each site independently. Where we used publications that pre-
viously compiled RSL reconstructions (e.g., refs. 37 and 45), the results were
used as presented in the compilation. RSL error was assumed to be a 2σ
range unless the original publication explicitly stated otherwise or if the
reconstruction was generated using a transfer function and a Random Mean
SE Standard Error of Prediction was reported, in which case this was treated
as a 1σrange. We did not reinterpret or reanalyze the published data, ex-
cept for the South American data (33, 59, 71, 90) that were mostly derived
from marine mollusks (vermetids). The radiocarbon ages for these data were
recalibrated using a more recent marine reservoir correction (91) and the
IntCal13 and MARINE13 radiocarbon age calibration curves (92).
Tide gauge records were drawn from the Permanent Service for Mean Sea
Level (PSMSL) (93, 94). We included all records that were either (i) longer
than 150 y, (ii) within 5 degrees distance of a proxy site and longer than 70 y,
or (iii) the nearest tide gauge to a proxy site that is longer than 20 y (Dataset
S1, b). We complement these with multicentury records from Amsterdam
(17001925 CE) (11), Kronstadt (17731993 CE) (95), and Stockholm (1774
2000 CE) (96), as compiled by PSMSL. Annual tide-gauge data were
smoothed by fitting a temporal GP model to each record and then
transforming the fitted model to decadal averages, both for computational
efficiency and because the decadal averages more accurately reflect the
recording capabilities of proxy records.
To incorporate information from a broader set of tide-gauge records, we
also included decadal averages from the Kalman smoother-estimated GSL for
18802010 CE of ref. 12. Off-diagonal elements of the GSL covariance matrix
were derived from an exponential decay function with a 3-y decorrelation
timescale. This timescale was set based on the mean temporal correlation
coefficient across all tide gauges using the annual PSMSL data, which ap-
proaches zero after 2 y.
Spatiotemporal Statistical Analysis. Hierarchical models (for a review tar-
geted at paleoclimatologists, see ref. 14) divide into different levels. The
hierarchical model we use separates into (i) a data level, which models
how the spatiotemporal sea-level field is recorded, with vertical and
temporal noise, by different proxies; (ii ) a process level, which models the
latent spatiotemporal field of RSL described by Eq. 1; and (iii ) a hyper-
parameter level. We used an empirical Bayesian analysis method, meaning
that, for computational efficiency, the hyperparameters used are point
estimates calibrated in a manner informed by the data (and described in
greater detail in Supporting Information); thus, our framework is called
an empirical hierarchical model. The output of the hierarchical model
includes a posterior probability distribution of the latent spatiotemporal
field fðx,tÞ, conditional on the point estimate hyperparameters. (Dataset
S3 provides the full time series and covariance of the posterior estimate of
GSL.) Our use of GP priors at the process level and normal likelihoods at
the data level renders the calculation of this conditional posterior ana-
lytically tractable (15).
At the data level, the observations yiare modeled as
yi=fðxi,tiÞ+wðxi,tiÞ+y0ðxiÞ+ey
i[2]
ti=
^
ti+et
i[3]
wðx,tÞGP0, σ2
wδðx,xÞδðt,tÞ[4]
y0ðxÞGP0, σ2
0δðx,xÞ[5]
where xiis the spatial location of observation i,tiis its age, wðx,tÞis a white
noise process that captures sea-level variability at a subdecadal level (which
we treat here as noise), ^
tiis the mean observed age, et
iand ey
iare errors in
the age and sea-level observations, y0ðxÞis a site-specific datum offset, and
δis the Kronecker delta function. The notation GPfμ,kðx,xÞÞg denotes a GP
with mean μand covariance function kðx,xÞ. For tide gauges, etis zero and
the distribution of eyis estimated during the GP smoothing process, in
which annual tide-gauge averages are assumed to have uncorrelated,
normally distributed noise with SD 3 mm. For proxy data, etand eyare
treated as independent and normally distributed, with an standard de-
viation (SD) specified for each data point based on the original publication.
Geochronological uncertainties are incorporated using the noisy input GP
method of ref. 97, which uses a first-order Taylor series approximation of the
latent process to translate errors in the independent variable into errors in
the dependent variable,
fðxi,tiÞfxi,^
ti+et
i
fxi,^
ti
t.[6]
The assumption that mean GSL over 100100 CE is equal to mean GSL
over 16001800 CE is implemented by conditioning on a set of pseudodata
with very broad uncertainties (SD of 100 m on each individual pseudodata
point) and a correlation structure that requires equality in the mean levels
over the two time windows.
At the process level, the GP priors for gðtÞ,lðxÞand mðx,tÞare given by
gðtÞGPn0, σ2
g0+σ2
gρt,t;τgo[7]
lðxÞGPICE5GðxÞ,σ2
lγðx,x;λlÞ[8]
mðx,tÞGP0, σ2
mγðx,x;λmÞρðt,t;τmÞ.[9]
Here, ICE5GðxÞdenotes the GIA rate given by the ICE5G-VM2-90 model of
ref. 98 for 17001950 CE. The temporal correlation function ρðt,t;τÞis a
Matérn correlation function with smoothness parameter 3/2 and scale τ.(The
choice of smoothness parameter 3/2 implies a functional form in which the
first temporal derivative is everywhere defined.) The spatial correlation
γðx,x;λÞis an exponential correlation function parameterized in terms of
the angular distance between xand x.
The hyperparameters of the model include the prior amplitudes σg0,
which is a global datum offset (for ML
2,1
, 118 mm); σg, which is the prior
amplitude of GSL variability (for ML
2,1
, 67 mm); σl, which is the prior SD of
slopes of the linear rate term (for ML
2,1
,1.1mm/y);andσm,whichisthe
prior amplitude of regional sea-level variability (for ML
2,1
,81mm).They
also include the timescales of global and regional variability, τgand τm(for
ML
2,1
, 136 y), the spatial scale of regional sea-level variability λm(for ML
2,1
,
7.7°), and the spatial scale of deviations of the linear term from the ICE5G-
VM2-90 GIA model, λl(for ML
2,1
,5.9°).IntheML
2,1
results presented in
the main text, it is assumed that τg=τm; four alternative sets of assump-
tions and calibrations of the hyperparameters are described in Supporting
Information.
Kopp et al. PNAS Early Edition
|
5of8
SUSTAINABILITY
SCIENCE
PNAS PLUS
Semiempirical Sea-Level Model. Our semiempirical sea-level model relates the
rate of GSL rise dh=dt to global mean temperature TðtÞ,
dh=dt =aðTðtÞT0ðtÞÞ +cðtÞ[10]
with
dT0ðtÞ=dt =ðTðtÞT0ðtÞÞ=τ1
dcðtÞ=dt =c=τ2,
where ais the sensitivity of the GSL rate to a deviation of TðtÞfrom an
equilibrium temperature T0ðtÞ,τ1is the timescale on which the actual tem-
perature relaxes toward the equilibrium temperature, and cis a temperature-
independent rate term with e-folding time τ2. The first term describes the GSL
response to climate change during the study period. The second term covers a
small residual trend arising from the long-term response to earlier climate
change (i.e., deglaciation), which is very slowly decaying over millennia and of
the order 0.1 mm/y in 2000 CE. It thus has a negligible effect on the modeled
GSL rise during the 20th and 21st centuries.
By comparison with Eq. 2, ref. 5 used the formulation
dh=dt =a1TðtÞT0,0+a2ðTðtÞT0ðtÞÞ +bðdT =dtÞ.[11]
The present model has two differences from that of ref. 5. First, we substitute
the temperature-independent term cðtÞfor term a1ðTðtÞT0,0Þand thus
eliminate the temperature dependence in this term. This modified term
describes a very slow component of sea-level adjustment that can capture
the tail end of the response to the last deglaciation. Second, we omit the
fast response term bðdT=dtÞbecause it is of no consequence on the long
timescales considered here.
We sample the posterior probability distribution of the parameter set
Ψ=fcðtÞ,T0ðtÞ,a,τ1,τ2g, specified as PðΨjgðtÞ,TðtÞÞ, using a Metropolis
Hastings (MH) algorithm (99) (Fig. S5Aand Dataset S1, j). The starting pa-
rameter set is a maximum-likelihood set, determined by simulated anneal-
ing. Sampled Markov Chains are thinned to every 500th sample, with the
first 1,000 samples discarded in a burn-in period.
We use two alternative temperature reconstructions (Fig. 1B): (i)the
global regularized expectation-maximization (RegEM) climate field re-
construction (CFR) temperature proxy of Mann et al. (1), incorporating the
HadCRUT3 instrumental data of ref. 100 after 1850 CE, and (ii ) the Marcott
et al. (2) RegEM global reconstruction. We use 11-y averages from the Mann
et al. reconstructions annual values, whereas the Marcott et al. reconstruction
reports 20-y average values. Because the number of proxy data in the Marcott
et al. reconstruction decreases toward present, we combine it with 20-y aver-
ages from the HadCRUT3 data (100) and align them over their period of overlap
(18501940 CE). The two temperature reconstructions are generally in good
agreement, although the Marcott et al. record shows 0.2 °C lower tempera-
tures before 1100 CE. This overall agreement provides confidence that the
true global temperature is represented within the uncertainties of the records,
whereas the modest differences motivate the use of both records to provide a
more realistic representation of uncertainty in the calculated GSL.
We denote the temperature reconstruction as SðtÞand treat it as noisy ob-
servations of TðtÞ;i.e.,ST,ΩÞ. To construct the temperature reconstruction
covariance Ω,weassumeSis an AR(1) time series, with variance as specified in the
reconstruction and a correlation e-folding time of 10 y. For each iteration iof our
MH algorithm, we draw n=100 samples Tjfrom TjS.WeassumethatSand Thave
uninformative priors, so that PðTjSÞ=PðSjTÞ. We then calculate the corresponding
sea-level time series hi,j=hðΨi,TjÞ, which we compare with the reconstructed GSL
gto calculate the posterior probability distribution PðΨijg,SÞ,
PðΨijg,SÞPðgjΨi,SÞPðΨijSÞ
=PðgjΨi,SÞPðΨiÞ
PðgjΨi,TÞPðTjSÞPðΨiÞ
1
nX
n
j=1
PgjΨi,TjPðΨiÞ
[12]
PgjΨi,Tj=j2πΣj1=2
×exp1
2^
ghi,jΣ1^
ghi,j[13]
where gNð^
g,ΣÞis taken from the ML
2,1
reconstruction. To calculate the
posterior distribution of the portion of GSL change explainable by the
semiempirical model, we simply take the distribution of hi,j. The prior and
posterior distributions of Ψare shown in Dataset S1,j.
To balance skill in modeling GSL with skill modeling the rate of change
of GSL, we taper the original covariance matrix Σrestimated from the
GP model. The resulting matrix Σ=Σr+λis the entrywise product of
the original matrix and an exponentially falling tapering function
λi,j=expðjtitjj=τcovÞ. We select a value of τcov =100 y, which is the
maximum-likelihood estimate for the Mann et al. calibration based on a
comparison of results for no tapering, a fully diagonal matrix, and values
of τcov f10, 50,100, 200, 500, 1,000gy(Fig. S5B).
Counterfactual hindcasts of 20th century GSL were calculated by substituting
T
jfor Tjin Eqs. 12 and 13, where each sample Tjwas transformed into Tj
such
that either (i) 20th century temperature followed a trend line fit to Tjfrom
500CEto1800CE,or(ii ) 20th century temperature was equal to the 5001800
CE average of Tj. The historical baseline Tj
was generated by replacing Tjafter
1900 CE with HadCRUT3, shifted so as to minimize the misfit between Had-
CRUT3 and Tjover 18501900. The nonanthropogenic fraction is calculated as
the distribution of hðΨi,Tj
ÞhðΨi,Tj
Þ(Fig. S4).
Projections of global mean temperature for the three RCPs were calculated
using the simple climate model MAGICC6 (101) in probabilistic mode, similar
to the approach of ref. 23. As described in ref. 102, the distribution of input
parameters for MAGICC6 was constructed through a Bayesian analysis based
upon historical observations (103, 104) and the equilibrium climate sensi-
tivity probability distribution of AR5 (105). We combined every set of pa-
rameters Ψiand historical temperatures Tjwith every single temperature
realization of MAGICC6, so the uncertainties are a combination of param-
eter uncertainty, initial condition uncertainty, and projected temperature
uncertainty.
ACKNOWLEDGMENTS. We thank M. Meinshausen for MAGICC6 tempera-
ture projections. We thank R. Chant, S. Engelhart, F. Simons, M. Tingley, and
two anonymous reviewers for helpful comments. This work was supported
by the US National Science Foundation (Grants ARC-1203414, ARC-1203415,
EAR-1402017, OCE-1458904, and OCE-1458921), the National Oceanic and
Atmospheric Administration (Grants NA11OAR431010 and NA14OAR4170085),
the New Jersey Sea Grant Consortium (publication NJSG-16-895), the Strategic
Environmental Resea rc h an d Development Program (Grant RC-2336), the
Natural Environmental Research Council (NERC; Grant NE/G003440/1), the
NERC Radiocarbon Facility, the Royal Society, and Harvard University. It is a
contribution to PALSEA2 (Palaeo-Constraints on Sea-Level Rise), which is a
working group of Past Global Changes/IMAGES (International Marine Past
Global Change Study) and an International Focus Group of the International
Union for Quaternary Research.
1. Mann ME, et al. (2009) Global signatures and dynamical origins of the Little Ice Age
and Medieval Climate Anomaly. Science 326(5957):12561260.
2. Marcott SA, Shakun JD, Clark PU, Mix AC (2013) A reconstruction of regional and global
temperature for the past 11,300 years. Science 339(6124):11981201.
3. PAGES 2K Consortium (2013) Continental-scale temperature variability during the
past two millennia. Nat Geosci 6(5):339346.
4. Jevrejeva S, Grinsted A, Moore J (2009) Anthropogenic forcing dominates sea level
rise since 1850. Geophys Res Lett 36(20):L20706.
5. Kemp AC, et al. (2011) Climate related sea-level variations over the past two mil-
lennia. Proc Natl Acad Sci USA 108(27):1101711022.
6. Lambeck K, Anzidei M, Antonioli F, Benini A, Esposito A (2004) Sea level in Roman
time in the Central Mediterranean and implications for recent change. Earth Planet
Sci Lett 224(3-4):563575.
7. Sivan D, et al. (2004) Ancient coastal wells of Caesarea Maritima, Israel, an in-
dicator for relative sea level changes during the last 2000 years. Earth Planet Sci
Lett 222(1):315330.
8. Woodroffe CD, McGregor HV, Lambeck K, Smithers SG, Fink D (2012) Mid-Pacific mic ro -
atolls record sea-level stability over the past 5000 yr. Geology 40(10):951954.
9. Lambeck K, Rouby H, Purcell A, Sun Y, Sambridge M (2014) Sea level and global ice
volumes from the Last Glacial Maximum to the Holocene. Proc Natl Acad Sci USA
111(43):1529615303.
10. Masson-Delmotte V, et al. (2013) Information from paleoclimate archives. Climate
Change 2013: The Physical Science Basis. Contribution of Working Group I to th e
Fifth Assessment Report of the Intergovernmental Panel on Climate Change,
eds Stocker TF, et al. (Cambridge Univ Press, Cambridge, UK), pp 383464.
11. van Veen J (1945) Bestaat er een geologische bodemdaling te amsterdam sedert
1700. Tijdschr K Ned Aardr Gen 62:236.
12. Hay CC, Morrow E, Kopp RE, Mitrovica JX (2015) Probabilistic reanalysis of twentieth-
century sea-level rise. Nature 517(7535):481484.
13. Cressie N, Wikle CK (2011) Statistics for Spatio-Temporal Data (Wiley, New York).
14. Tingley MP, et al. (2012) Piecing together the past: Statistical insights into paleo-
climatic reconstructions. Quat Sci Rev 35:122.
6of8
|
www.pnas.org/cgi/doi/10.1073/pnas.1517056113 Kopp et al.
15. Rasmussen C, Williams C (2006) Gaussian Processes for Machine Learning (MIT Press,
Cambridge, MA).
16. Yin J, Goddard PB (2013) Oceanic control of sea level rise patterns along the east
coast of the United States. Geophys Res Lett 40(20):55145520.
17. Mitrovica JX, et al. (2011) On the robustness of predictions of sea level fingerprints.
Geophys J Int 187(2):729742.
18. Kopp RE, Hay CC, Little CM, Mitrovica JX (2015) Geographic variability of sea-level
change. Curr Clim Change Rep 1(3):192204.
19. Woodworth PL, Maqueda MÁM, Roussenov VM, Williams RG, Hughes CW (2014)
Mean sea-level variability along the northeast American Atlantic coast and the roles
of the wind and the overturning circulation. J Geophys Res 119(12):89168935.
20. Grinsted A, Moore JC, Jevrejeva S (2009) Reconstructing sea level from paleo and
projected temperatures 200 to 2100 AD. Clim Dyn 34(4):461472.
21. Rahmstorf S (2007) A semi-empirical approach to projecting future sea-level rise.
Science 315(5810):368370.
22. Vermeer M, Rahmstorf S (2009) Global sea level linked to global temperature. Proc
Natl Acad Sci USA 106(51):2152721532.
23. Schaeffer M, Hare W, Rahmstorf S, Vermeer M (2012) Long-term sea-level rise im-
plied by 1.5 °C and 2 °C warming levels. Nat Clim Change 2:867870.
24. Bittermann K, Rahmstorf S, Perrette M, Vermeer M (2013) Predictability of twentieth
century sea-level rise from past data. Environ Res Lett 8(1):014013.
25. Gregory JM, et al. (2013) Twentieth-century global-mean sea level rise: Is the whole
greater than the sum of the parts? J Clim 26(13):44764499.
26. Becker M, Karpytchev M, Lennartz-Sassinek S (2014) Long-term sea level trends:
natural or anthropogenic? Geophys Res Lett 41(15):55715580.
27. Dangendorf S, et al. (2015) Detecting anthropogenic footprints in sea level rise. Nat
Commun 6:7849.
28. Meehl G, et al. (2007) Global climate projections. Climate Change 2007: The Physical
Science Basis: Fourth Assessment Report of the Intergovernmental Panel on Climate
Change, eds Solomon S, et al. (Cambridge Univ Press, Cambridge, UK), pp 747845.
29. Church JA, et al. (2013) Sea level change. Climate Change 2013: The Physical Science
Basis, eds Stocker TF, et al. (Cambridge Univ Press, Cambridge, UK), pp 11371216.
30. Kopp RE, et al. (2014) Probabilistic 21st and 22nd century sea-level projections at a
global network of tide gauge sites. Earths Future 2(8):383406.
31. Horton BP, Rahmstorf S, Engelhart SE, Kemp AC (2014) Expert assessment of sea-
level rise by AD 2100 and AD 2300. Quat Sci Rev 84:16.
32. Joughin I, Smith BE, Medley B (2014) Marine ice sheet collapse potentially under way
for the Thwaites Glacier Basin, West Antarctica. Science 344(6185):735738.
33. Angulo RJ, Giannini PCF, Suguio K, Pessenda LCR (1999) Relative sea-level changes in
the last 5500 years in southern Brazil (LagunaImbituba region, Santa Catarina
State) based on vermetid
14
C ages. Mar Geol 159(1-4):323339.
34. Barlow NLM, et al. (2014) Salt-marsh reconstructions of relative sea-level change in
the North Atlantic during the last 2000 years. Quat Sci Rev 99:116.
35. Baxter AJ (1997) Late Quaternary palaeoenvironments of the Sandveld, Western
Cape Province, South Africa. Ph.D. thesis (University of Cape Town, Cape Town,
South Africa).
36. Cinquemani LJ, Newman WS, Sperling JA, Marcus LF, Pardi RR (1982) Holocene sea
level changes and vertical movements along the east coast of the United States:
A preliminary report. Holocene Sea Level Fluctuations, Magnitude and Causes,
ed Colquhoun D (Univ South Carolina, Columbia), pp 1333.
37. Compton JS (2001) Holocene sea-level fluctuations inferred from the evolution of
depositional environments of the southern Langebaan Lagoon salt marsh, South
Africa. Holocene 11(4):395405.
38. Dawson S, Smith D (1997) Holocene relative sea-level changes on the margin of a
glacio-isostatically uplifted area: An example from northern Caithness, Scotland.
Holocene 7(1):5977.
39. Deevey ES, Gralenski LJ, Hoffren V (1959) Yale natural radiocarbon measurements
IV. Radiocarbon 1:144159.
40. Delibras C, Laborel J (1969) Recent variations of the sea level along the Brazilian
coast. Quaternaria 14:4549.
41. Donnelly JP, et al. (2001) Sedimentary evidence of intense hurricane strikes from
New Jersey. Geology 29(7):615618.
42. Donnelly JP, Cleary P, Newby P, Ettinger R (2004) Coupling instrumental and geo-
logical records of sea-level change: Evidence from southern New England of an in-
crease in the rate of sea-level rise in the late 19th century. Geophys Res Lett 31(5):
L05203.
43. Donnelly JP, Butler J, Roll S, Wengren M, Webb T (2004) A backbarrier overwash
record of intense storms from Brigantine, New Jersey. Mar Geol 210:107121.
44. Donnelly JP (2006) A revised late Holocene sea-level record for northern Mas-
sachusetts, USA. J Coast Res 22(5):10511061.
45. Engelhart SE, Horton BP (2012) Holocene sea level database for the Atlantic coast of
the United States. Quat Sci Rev 54:1225.
46. Garcia-Artola A, Cearreta A, Leorri E, Irabien M, Blake W (2009) Coastal salt-marshes
as geological archives of recent sea-level changes. Geogaceta 47:109112.
47. Gehrels WR, et al. (2005) Onset of recent rapid sea-level rise in the western Atlantic
Ocean. Quat Sci Rev 24:20832100.
48. Gehrels WR, et al. (2006) Late Holocene sea-level changes and isostasy in western
Denmark. Quat Res 66(2):288302.
49. Gehrels WR, et al. (2006) Rapid sea-level rise in the North Atlantic Ocean since the
first half of the nineteenth century. Holocene 16(7):949965.
50. Gehrels WR, Hayward B, Newnham RM, Southall KE (2008) A 20th century acceler-
ation of sea-level rise in New Zealand. Geophys Res Lett 35(2):L02717.
51. Gehrels WR, Dawson DA, Shaw J, Marshall WA (2011) Using Holocene relative sea-
level data to inform future sea-level predictions: An example from southwest Eng-
land. Global Planet Change 78(3-4):116126.
52. Gehrels WR, et al. (2012) Nineteenth and twentieth century sea-level changes in
Tasmania and New Zealand. Earth Planet Sci Lett 315-316:94102.
53. Gehrels WR, Anderson WP (2014) Reconstructing Holocene sea-level change from
coastal freshwater peat: A combined empirical and model-based approach. Mar
Geol 353:140152.
54. Gibb J (1986) A New Zealand regional Holocene eustatic sea-level curve and its appli-
cation to determination of vertical tectonic movement. RSocNZBull24:377395.
55. González JL, Törnqvist TE (2009) A new Late Holocene sea-level record from the Mis-
sissippi Delta: Evidence for a climate/sea level connection? Quat Sci Rev 28:17371749.
56. Goodwin ID, Harvey N (2008) Subtropical sea-level history from coral microatolls in
the Southern Cook Islands, since 300 AD. Mar Geol 253(1-2):1425.
57. Horton BP, et al. (2009) Holocene sea-level changes along the North Carolina
coastline and their implications for glacial isostatic adjustment models. Quat Sci Rev
28:17251736.
58. Horton BP, et al. (2013) Influence of tidal-range change and sediment compaction on
holocene relative sea-level change in New Jersey, USA. J Quaternary Sci 28(4):403411.
59. Ireland S (1988) Holocene coastal changes in Rio de Janeiro State, Brazil. Doctoral
thesis (Durham University, Durham, UK).
60. Kemp AC, et al. (2013) Sea-level change during the last 2500 years in New Jersey,
USA. Quat Sci Rev 81:90104.
61. Kemp AC, et al. (2014) Late Holocene sea- and land-level change on the U.S.
southeastern Atlantic coast. Mar Geol 357:90100.
62. Kemp AC, et al. (2015) Relative sea-level change in Connecticut (USA) during the last
2200 yrs. Earth Planet Sci Lett 428:217229.
63. Leorri E, Horton BP, Cearreta A (2008) Development of a foraminifera-based transfer
function in the Basque marshes, N. Spain: Implications for sea-level studies in the Bay
of Biscay. Mar Geol 251(1-2):6074.
64. Leorri E, Cearreta A (2009) Anthropocene versus Holocene relative sea-level rise rates
in the southern Bay of Biscay. Geogaceta 46:127130.
65. Leorri E, Cearreta A, Milne G (2012) Field observations and modelling of Holocene
sea-level changes in the southern Bay of Biscay: Implication for understanding cur-
rent rates of relative sea-level change and vertical land motion along the Atlantic
coast of SW Europe. Quat Sci Rev 42:5973.
66. Long A, Tooley M (1995) Holocene sea-level and crustal movements in Hampshire
and Southeast England, United Kingdom. J Coast Res 17:299310.
67. Long A, Scaife R, Edwards R (2000) Stratigraphic architecture, relative sea-level, and
models of estuary development in southern England: New data from Southampton
Water. Geol Soc Lond Spec Publ 175:253279.
68. Long AJ, et al. (2012) Relative sea-levelchange in Greenland during thelast 700 yrs and
ice sheet response to the Little Ice Age. Earth Planet Sci Lett 315-316:7685.
69. Long AJ, et al. (2014) Contrasting records of sea-level change in the eastern and
western North Atlantic during the last 300 years. Earth Planet Sci Lett 388:110122.
70. Martin A (1968) Pollen analysis of Groenvlei lake sediments, Knysna (South Africa).
Rev Palaeobot Palynol 7(2):107144.
71. Martin L, Suguio K (1978) Excursion route along the coastline between the town of
Cananéia (State of São Paulo) and Guaratiba outlet (State of Rio de Janeiro). In-
ternational Symposium on Global Changes in South America during the Quaternary
(Univ Sao Paolo, Sao Paolo, Brazil), Vol 2, pp 264274.
72. Miller KG, et al. (2009) Sea-level rise in New Jersey over the past 5000 years: Impli-
cations to anthropogenic changes. Global Planet Change 66(1-2):1018.
73. Nydick KR, Bidwell AB, Thomas E, Varekamp JC (1995) A sea-level rise curve from
Guilford, Connecticut, USA. Mar Geol 124(1-4):137159.
74. Pardi RR, Tomecek L, Newman WS (1984) Queens College radiocarbon measure-
ments IV. Radiocarbon 26(3):412430.
75. Redfield AC, Rubin M (1962) The age of salt marsh peat and its relation to recent
changes in sea level at Barnstable, Massachusetts. Proc Natl Acad Sci USA 48(10):
17281735.
76. Saher MH, et al. (2015) Sea-level changes in Iceland and the influence of the North
Atlantic Oscillation during the last half millennium. Quat Sci Rev 108:2336.
77. Scott DB, Gayes PT, Collins ES (1995) Mid-Holocene precedent for a future rise in sea
level along the Atlantic Coast of North America. J Coast Res 11(3):615622.
78. Shennan I, Horton B (2002) Relative sea-level changes and crustal movements of the
UK. J Quaternary Sci 16:511526.
79. Spaur C, Snyder S (1999) Coastal wetlands evolution at the leading edge of the
marine transgression: Jarrett Bay, North Carolina. J N C Acad Sci 115(1):2046.
80. Stéphan P, et al. (2015) Holocene salt-marsh sedimentary infilling and relative sea-
level changes in West Brittany (France) using foraminifera-based transfer functions.
Boreas 44(1):153177.
81. Strachan K, Finch J, Hill T, Barnett R (2014) A late Holocene sea-level curve for the
east coast of South Africa. S Afr J Sci 110(1/2):2013-0198.
82. Stuiver M, Daddario JJ (1963) Submergence of the New Jersey Coast. Science
142(3594):951.
83. Stuiver M, Deevey ES, Rouse I (1963) Yale natural radiocarbon measurements.
Radiocarbon 5:312341.
84. Szkornik K, Gehrels WR, Murray AS (2008) Aeolian sand movement and relative sea-level
rise in Ho Bugt, western Denmark, during the Little Ice Age.Holocene 18(6):951965.
85. Van de Plassche O (1991) Late Holocene sea-level fluctuations on the shore of
Connecticut inferred from transgressive and regressive overlap boundaries in salt-
marsh deposits. J Coast Res 11:159179.
86. Van de Plassche O, van der Borg K, de Jong AF (1998) Sea levelclimate correlation
during the past 1400 yr. Geology 26(4):319322.
87. Van de Plassche O, Van der Borg K, De Jong A (2002) Relative sea-level rise across the
Eastern Border fault (Branford, Connecticut): Evidence against seismotectonic
movements. Mar Geol 184(1-2):6168.
Kopp et al. PNAS Early Edition
|
7of8
SUSTAINABILITY
SCIENCE
PNAS PLUS
88. Woodroffe SA, Long AJ, Milne GA, Bryant CL, Thomas AL (2015) New constraints
on late Holocene eustatic sea-level changes from Mahé, Seychelles. Quat Sci Rev
115:116.
89. McGregor H, Fischer M, Gagan M, Fink D, Woodroffe C (2011) Environmental control
of the oxygen isotope composition of Porites coral microatolls. Geochim Cosmochim
Acta 75(14):39303944.
90. Milne GA, Long AJ, Bassett SE (2005) Modelling Holocene relative sea-level obser-
vations from the Caribbean and South America. Quat Sci Rev 24:11831202.
91. Angulo RJ, de Souza MC, Reimer PJ, Sasaoka SK (2005) Reservoir effect of the
southern and southeastern Brazilian coast. Radiocarbon 47(1):6773.
92. Reimer PJ, et al. (2013) IntCal13 and Marine13 radiocarbon age calibration curves
050,000 years cal BP. Radiocarbon 55(4):18691887.
93. Holgate SJ, et al. (2013) New data systems and products at the Permanent Service for
Mean Sea Level. J Coast Res 29(3):493504.
94. Permanent Service for Mean Sea Level (PSMSL) (2014) Tide gauge data. Available at
www.psmsl.org/data/obtaining/. Accessed September 29, 2014.
95. Bogdanov VI, Laitos G (2000) Mean Monthly Series of Sea Level Observations (1777
1993) at the Kronstadt Gauge (Finnish Geodetic Inst, Kirkkonummi, Finland).
96. Ekman M (1988) The worlds longest continued series of sea level observations. Pure
Appl Geophys 127(1):7377.
97. McHutchon A, Rasmussen C (2011) Gaussian process training with input noise.
Advances in Neural Information Processing Systems (Neural Inf Process Syst Found,
La Jolla, CA), Vol 24, pp 13411349.
98. Peltier WR (2004) Global glacial isostasy and the surface of the ice-age
Earth: The ICE-5G (VM2) model and GRACE. Annu Rev Earth Planet Sci 32:
111149.
99. Hastings WK (1970) Monte Carlo sampling methods using Markov chains and their
applications. Biometrika 57(1):97109.
100. Brohan P, Kennedy JJ, Harris I, Tett SFB, Jones PD (2006) Uncertainty estimates in
regional and global observed temperature changes: A new data set from 1850.
J Geophys Res 111(D12):D12106.
101. Meinshausen M, Raper SCB, Wigley TML (2011) Emulating coupled atmosphere-
ocean and carbon cycle models with a simpler model, MAGICC6 Part 1: Model
description and calibration. Atmos Chem Phys 11(4):14171456.
102. Rasmussen DM, Kopp RE (2015) Appendix A. Physical climate projections. Economic
Risks of Climate Change: An American Prospectus, eds Houser T, Hsiang S, Kopp R,
Larsen K (Columbia Univ Press, New York), pp 219248.
103. Meinshausen M, et al. (2009) Greenhouse-gas emission targets for limiting global
warming to 2 °C. Nature 458(7242):11581162.
104. Rogelj J , Meinshaus en M, Kn utti R (2012) G lobal warmi ng under old an d new
scenarios using IPCC climate sensitivity range estimates. Nat Clim Change 2:
248253.
105. Collins M, et al. (2013) Long-term climate change: Projections, commitments and
irreversibility. Climate Change 2013: The Physical Science Basis, eds Stocker TF, et al.
(Cambridge Univ Press, Cambridge, UK), pp 10291136.
106. Church J, White N (2011) Sea-level rise from the late 19th to the early 21st century.
Surv Geophys 32(4):585602.
8of8
|
www.pnas.org/cgi/doi/10.1073/pnas.1517056113 Kopp et al.
Supporting Information
Kopp et al. 10.1073/pnas.1517056113
Sensitivity Tests for Reconstruction
We consider five alternative empirical calibrations of the
hyperparameters Θ=fσg,σl,σm,σw,σ0,σg0,τg,τm,λl,λmg(Data-
set S1, c, and Fig. S3). (i) For prior ML
2,2
,Θis empirically
calibrated through a hybrid global simulated annealing/local se-
quential quadratic programming optimization to find the
hyperparameters that maximize the likelihood of the model
conditional upon the observations. (ii)PriorML
2,1
is similar
to prior ML
2,2
, but with the constraint that τm=τg, i.e., that
there is one timescale hyperparameter for the nonlinear
terms. (iii) Prior ML
1,1
is similar to prior ML
2,2
, but with the
constraints that τm=τgand σm=σg, i.e., that there is one time-
scale hyperparameter and one amplitude hyperparameter for the
nonlinear terms. (iv) For prior NC, σgand τgare optimized to
maximize the likelihood of a GP model fit to the curve derived
from the detrended North Carolina RSL reconstruction. The re-
maining hyperparameters are then empirically optimized to maxi-
mize the likelihood of the model conditional upon the observations.
(v)ForpriorGr,σgand τgare optimized to maximize the likelihood
of a GP model fit to the ref. 20 GSL hindcast. The remaining
hyperparameters are then empirically optimized to maximize the
likelihood of the model conditional upon the observations.
In all cases, the parameters were optimized conditional only
upon observations with 2σage uncertainties of less than ±50
years. In order to ensure that the optimization process does not
confuse the interpretations of the terms reflecting the processes
of interest and the white-noise term, we optimized under the
constraint that τ
g
and τ
m
>100 y. To ensure that this is a rea-
sonable constraint, we also tested a model in which the amplitude
of the global term, σ
g
, was set equal to zero; thus, in this model,
all observations were interpreted as reflecting regional changes.
Without the inclusion of the global term, the optimal value of τ
m
is 123 y, consistent with the constraint.
From an interpretive perspective, the main difference among
the calibrated prior is the timescale of GSL variability τg(Dataset
S1, c). In priors ML
2,1
and ML
1,1
, this timescale is 100 y, whereas,
for ML
2,2
,itis2,000 y, yielding an overly smooth GSL curve; other
priors have intermediate timescales. Shorter timescales lead a greater
proportion of centennial-scale variability to be attributed to
GSL. Although ML
2,2
has the highest marginal likelihood, we
focus on ML
2,1
in presenting results, as inspection of within-20th
century rate estimates for ML
2,2
suggests oversmoothing. Prior
estimates of rates and variability are shown in Dataset S1, d.
We also consider the application of the model with the ML
2,1
before different subsets of data (Dataset S1, f). The decline in
GSL between 1000 CE and 1400 CE is robust to the removal of
North Atlantic data (P=0.91 in subset -NAtlantic+GSL) and to
the consideration only of data from the Atlantic and the Medi-
terranean (P=1.00 in subset +AtlanticMediterranean+GSL). It
is, however, not robust to the consideration only of data from
South America (P=0.48 in subset +SAmerica+GSL) or only
of data outside the Atlantic and Mediterranean (P=0.25 in
subset -AtlanticMediterranean+GSL).
To assess the impact of the exogenous ref. 12 GSL curve for the
20th century, we consider each subset of data with and without the
inclusion of this curve (indicated in Dataset S1, f by +GSL and
-GSL). Twentieth century rates have broader errors without
the exogenous 20th century curve but are in general agreement
with that curve (e.g., +All+GSL: 1.38 ±0.15 mm/y; +All-GSL:
1.31 ±0.33 mm/y).
a) Proxy data
b) Tide-gauge data
Fig. S1. Locations of sites with (A) proxy data and (B) tide-gauge data included in the analysis.
Kopp et al. www.pnas.org/cgi/content/short/1517056113 1of6
Fig. S2. Model fits under prior ML
2,1
at eight illustrative sites: (A) East River Marsh, CT; (B) Sand Point, NC, (C) Vioarholmi, Iceland, (D) Loch Laxford, Scotland,
(E) Sissimut, Greenland, (F) Caesarea, Israel, (G) Christmas Island, Kiribati, and (H) Kariega Estuary, South Africa. (Note that the model fit at each site is informed
by all observations, not just those at the illustrated site.) Red boxes show all data points within 0.1 degrees of the centroid of the named site. Errors are ±2σ.
Kopp et al. www.pnas.org/cgi/content/short/1517056113 2of6
Fig. S3. (AE) GSL estimates under priors (A)ML
2,1
,(B)ML
2,2
,(C)ML
1,1
,(D) Gr, and (E) NC. (F) GSL reconstruction under prior ML
2,1
(black) compared with the
hindcast of ref. 20 (G09 hindcast; red), a curve derived from the North Carolina RSL curve by detrending and the addition of ±10 mm (2σ) errors after ref. 5 (NC
pseudo-GSL; green), and the 19th20th century GMSL reconstructions of ref. 12 (Hay2015; magenta) and ref. 106 (CW2011; yellow). Multicentury records are
aligned relative to the 16001800 CE average; 19th20th century reconstructions are aligned to match our GSL reconstruction in 2000 CE. Errors are ±1σ.
Kopp et al. www.pnas.org/cgi/content/short/1517056113 3of6
-5
0
5
10
15
Temp. anomaly
1800 1850 1900 1950 2000
-0.6
-0.4
-0.2
0
0.2
0.4
0.6
Temp. anomaly
1800 1850 1900 1950 2000
-0.6
-0.4
-0.2
0
0.2
0.4
0.6
1800 1850 1900 1950 2000
-5
0
5
10
15
Temp. anomaly
1800 1850 1900 1950 2000
-0.6
-0.4
-0.2
0
0.2
0.4
0.6
1800 1850 1900 1950 2000
Temp. anomaly
1800 1850 1900 1950 2000
-0.6
-0.4
-0.2
0
0.2
0.4
0.6
Mann et al. 2009 Marcott et al. 2013
GSL (cm above 1900 CE)
Fig. S4. Counterfactual hindcasts of global mean sea-level rise in the absence of anthropogenic warming. Each row assumes a different counterfactual
temperature scenario (see Materials and Methods), and each column represents model calibration to a different temperature reconstruction (Inset). In the
temperature Insets, the black lines represent the original temperature reconstruction to 1900, the blue line represents the counterfactual scenario, and the red
line represents the HadCRUT3 temperature reconstruction for the 20th century. In the main plots, the blue and red curves correspond, respectively, tothe
HadCRUT3 and counterfactual temperature scenarios. The difference between them can be interpreted as the anthropogenic GSL rise. Heavy shading, 67%
credible interval; light shading, 90% credible interval.
Kopp et al. www.pnas.org/cgi/content/short/1517056113 4of6
a
Probability
0.4 0.6 0.8 1
0
0.05
0.1
0.15
τ
200 400 600 800
0
0.1
0.2
0.3
τc
0.5 1 1.5
0
0.05
0.1
0.15
× 104
T0(501)
-0.2 0 0.2
0
0.05
0.1
0.15
0.2
T0(-2000)
-0.5 0 0.5 1
0
0.02
0.04
0.06
c(2000)
Probability
0 0.02 0.04
0
0.1
0.2
Marcott et. al PosteriorMann et. al PosteriorParameter Prior
010 50 100 200 500 1000 NaN
10−35
10−30
10−25
10−20
10−15
10−10
10−5
100
τcov
Likelihood SL
a
b
Fig. S5. (A) Probability distributions of semiempirical model parameters. Prior distributions (gray) are not shown in full where scaling axes to display them
would render posteriors obscure. (B) Likelihood of different values of τcov . Black dots show likelihoods across all semiempirical projections, and the bars show
the 5th/17th/83rd/95th percentile conditional likelihoods for individual semiempirical projections. A value of τcov =100 y is used in the analysis. Orange [green]
represents calibration with the Mann et al. (1) [Marcott et al. (2)] temperature curve.
Kopp et al. www.pnas.org/cgi/content/short/1517056113 5of6
Dataset S1. (a) A summary of the sites included in the Common Era database, (b) a summary of the tide gauges incorporated into the
metaanalysis, (c) hyperparameters of the different priors for the empirical hierarchical model, (d) prior estimates of GSL rates and
amplitude of variability under different priors, (e) posterior estimates of GSL rates and amplitude of variability under different priors, (f)
GSL rates under different data subsets, (g) RSL rates at different sites, (h) semiempirical estimates of the probability that the observed
GSL rise exceeds counterfactual projections, (i) semiempirical GSL projections, and (j) the distribution of semiempirical model parameters
Dataset S1
Dataset S2. The full Common Era RSL reconstruction database
Dataset S2
Dataset S3. Time series and covariance matrices of GSL under the five different priors
Dataset S3
Kopp et al. www.pnas.org/cgi/content/short/1517056113 6of6
ii“CEGSL4b” — 2016/1/20 — 10:29 — page 1 — #18 i
i
iii
i
Dataset S1
Kopp et al.
Dataset S1a. Proxy sites and original studies used in this analysis
Location Mean Latitude (N) Mean Longitude (E) Median Age Range (CE) N References
Western North Atlantic Ocean
Connecticut, USA 41.3 -71.9 -1174 to 1981 162 ref. 36, 42, 62, 73,
85–87
Florida, USA 30.6 -81.7 -559 to 1999 65 ref. 61
Greenland 65.7 -51.0 1328 to 1950 44 ref. 68
Louisiana, USA 29.9 -91.8 448 to 1807 23 ref. 55
Massachusetts, USA 42.2 -70.8 -1440 to 1963 39 ref. 5, 44, 45, 75,
83
New Jersey, USA 39.3 -74.6 -476 to 2000 151 ref. 36, 41, 43, 58,
60, 72, 74, 82
North Carolina, USA 35.6 -75.9 -1694 to 2000 180 ref. 5, 36, 57, 79
Nova Scotia, Canada 44.7 -63.3 -836 to 1996 78 ref. 47, 77
Eastern North Atlantic Ocean
Brittany, France 48.3 -4.3 -872 to 1467 7 ref. 80
Denmark 55.6 -8.3 -808 to 1881 22 ref. 48, 84
Iceland 64.8 -22.4 -141 to 2002 93 ref. 49, 76
Isle of Wight, UK 50.7 -1.4 -517 to 2011 26 ref. 66, 67, 69
Scotland, UK 58.4 -4.8 -198 to 2009 108 ref. 34, 38, 78
South West England, UK 50.3 -3.9 -482 to 1448 11 ref. 51, 53
Spain 43.4 -2.9 -1075 to 2003 52 ref. 46, 63–65
South Atlantic Ocean
Rio de Janeiro, Brazil -23.0 -43.5 -1302 to 1388 16 ref. 40, 59, 71
Santa Catarina, Brazil -28.5 -48.8 -688 to 1834 20 ref. 33
South Africa -33.7 26.7 -1197 to 2010 41 ref. 35, 37, 39, 70,
81
Mediterranean Sea
Israel 32.5 34.9 -10 to 1132 65 ref. 7
Pacific Ocean
Christmas Island, Kiribati 2.0 -157.5 -1050 to 1860 72 ref. 8, 89
Cook Islands -20.2 -159.8 363 to 1900 12 ref. 56
New Zealand -46.5 169.7 -299 to 1991 15 ref. 50, 54
Tasmania, Australia -42.3 147.9 1820 to 2004 28 ref. 52
Indian Ocean
Seychelles -4.7 55.5 -646 to 1442 10 ref. 88
Mean latitude, mean longitude and the range of median data point age estimates are shown for the N individual data points from each
site.
Kopp et al. Dataset 1 1
ii“CEGSL4b” — 2016/1/20 — 10:29 — page 2 — #19 i
i
iii
i
Dataset S1b. Tide gauges used in this analysis
Name Latitude (N) Longitude (E) First Year Last Year PSMSL ID
AMSTERDAM 52.37 4.90 1700 1925 —
KRONSTADT 59.98 29.77 1777 1993 —
STOCKHOLM 59.32 18.08 1774 2000 —
BREST 48.38 -4.49 1807 2013 1
SWINOUJSCIE 53.92 14.23 1811 1999 2
SHEERNESS 51.45 0.74 1833 2006 3
CUXHAVEN 2 53.87 8.72 1843 2010 7
WISMAR 2 53.90 11.46 1849 2012 8
MAASSLUIS 51.92 4.25 1848 2013 9
SAN FRANCISCO 37.81 -122.47 1855 2013 10
WARNEMUNDE 2 54.17 12.10 1856 2012 11
NEW YORK 40.70 -74.01 1856 2013 12
LIVERPOOL 53.40 -3.00 1858 1983 15
VLISSINGEN 51.44 3.60 1862 2013 20
ABERDEEN II 57.15 -2.08 1862 1965 21
HOEK VAN HOLLAND 51.98 4.12 1864 2013 22
DEN HELDER 52.96 4.75 1865 2013 23
HARLINGEN 53.18 5.41 1865 2013 25
IJMUIDEN 52.46 4.55 1872 2013 32
STAVANGER 58.97 5.73 1919 2012 47
BERGEN 60.40 5.32 1916 2012 58
NORTH SHIELDS 55.01 -1.44 1896 2013 95
HALIFAX 44.67 -63.58 1896 2011 96
FERNANDINA BEACH 30.67 -81.47 1898 2013 112
TROIS-RIVIERES 46.33 -72.55 1925 2013 126
PHILADELPHIA 39.93 -75.14 1901 2013 135
DUNEDIN II -45.88 170.51 1900 2013 136
BALTIMORE 39.27 -76.58 1903 2013 148
GALVESTON II 29.31 -94.79 1909 2013 161
ATLANTIC CITY 39.35 -74.42 1912 2013 180
PORTLAND 43.66 -70.25 1912 2013 183
NEWLYN 50.10 -5.54 1916 2013 202
CHARLESTON I 32.78 -79.92 1922 2013 234
BOSTON 42.35 -71.05 1921 2013 235
WEST-TERSCHELLING 53.36 5.22 1921 2013 236
PENSACOLA 30.40 -87.21 1924 2011 246
PORT SAID 31.25 32.30 1923 1946 253
SEWELLS POINT 36.95 -76.33 1928 2013 299
ANNAPOLIS 38.98 -76.48 1929 2013 311
PORTSMOUTH 50.80 -1.11 1962 2012 350
NEWPORT 41.51 -71.33 1931 2013 351
WASHINGTON DC 38.87 -77.02 1931 2013 360
SANDY HOOK 40.47 -74.01 1933 2013 366
WOODS HOLE 41.52 -70.67 1933 2013 367
FORT PULASKI 32.03 -80.90 1935 2013 395
WILMINGTON 34.23 -77.95 1936 2013 396
NEW LONDON 41.36 -72.09 1939 2013 429
EUGENE ISLAND 29.37 -91.39 1940 1974 440
ST JEAN DE LUZ 43.40 -1.68 1943 2010 469
SANTANDER I 43.46 -3.79 1944 2012 485
REYKJAVIK 64.15 -21.94 1957 2013 638
CANANEIA -25.02 -47.93 1955 2004 726
PORT ELIZABETH -33.95 25.63 1980 2010 820
SIMONS BAY -34.19 18.44 1958 2013 826
MALIN HEAD 55.37 -7.33 1959 2001 916
KNYSNA -34.05 23.05 1961 2013 950
DEVONPORT 50.37 -4.19 1962 2013 982
ILHA FISCAL -22.90 -43.17 1965 2013 1032
BRIDGEPORT 41.17 -73.18 1965 2013 1068
WICK 58.44 -3.09 1965 2012 1109
ULLAPOOL 57.90 -5.16 1983 2013 1112
CAPE MAY 38.97 -74.96 1966 2013 1153
SPRING BAY -42.55 147.93 1992 2013 1216
CHRISTMAS ISLAND II 1.98 -157.48 1974 2010 1371
RAROTONGA -21.20 -159.77 1978 2001 1453
DUCK PIER OUTSIDE 36.18 -75.75 1985 2013 1636
Kopp et al. Dataset 1 2
ii“CEGSL4b” — 2016/1/20 — 10:29 — page 3 — #20 i
i
iii
i
Dataset S1c. Hyperparameters under dierent modeling assumptions
Model log likelihood ggllmmmw0g0
(mm) (yr) (mm/yr) ()(mm) (yr) ()(mm) (mm) (mm)
ML2,1-12260 57.3 100.0 1.1 5.5 50.9 — 6.4 20.9 18.1 121.1
ML2,2-12257 772.3 2004.9 0.9 2.7 57.0 100.0 9.0 20.9 16.6 0.2
ML1,1-12266 53.7 100.0 1.1 5.7 7.0 20.8 18.2 118.5
NC -12255 396.9 692.3 0.9 1.9 57.7 100.0 7.4 20.8 0.4 235.9
Gr -12259 165.8 311.7 1.1 4.9 53.1 100.0 7.1 21.0 18.1 121.2
Dataset S1d. Prior estimates of rates of GSL change under dierent modeling assumptions (mm/yr)
ML2,1ML2,2ML1,1NC Gr
0–300 0.00 ±0.48 0.00 ±1.03 0.00 ±0.45 0.00 ±1.50 0.00 ±1.04
300–700 0.00 ±0.40 0.00 ±0.75 0.00 ±0.38 0.00 ±1.24 0.00 ±0.91
700–1000 0.00 ±0.53 0.00 ±0.70 0.00 ±0.50 0.00 ±1.37 0.00 ±1.09
1000–1400 0.00 ±0.40 0.00 ±0.75 0.00 ±0.38 0.00 ±1.24 0.00 ±0.91
1400–1800 0.00 ±0.38 0.00 ±1.03 0.00 ±0.36 0.00 ±1.41 0.00 ±0.92
1800–1900 0.00 ±1.11 0.00 ±1.26 0.00 ±1.04 0.00 ±1.81 0.00 ±1.48
1860–1900 0.00 ±1.56 0.00 ±1.29 0.00 ±1.46 0.00 ±1.89 0.00 ±1.65
1900–2000 0.00 ±1.16 0.00 ±1.28 0.00 ±1.08 0.00 ±1.78 0.00 ±1.47
Amplitude (cm) ±10 (815)±17 (837)±11 (814)±27 (1448)±20 (1233)
Errors are ±2. Amplitude row indicates the median (5th–95th percentile) prior estimate of the amplitude of variability over 0–1900 CE.
Dataset S1e. Posterior estimates of rates of GSL change under dierent modeling assumptions (mm/yr)
ML2,1ML2,2ML1,1NC Gr
0–300 0.13 ±0.25 (0.87) 0.18 ±0.21 (0.96) 0.14 ±0.25 (0.88) 0.17 ±0.26 (0.91) 0.17 ±0.25 (0.92)
300–700 0.08 ±0.20 (0.79) 0.04 ±0.18 (0.68) 0.07 ±0.20 (0.78) 0.05 ±0.21 (0.68) 0.06 ±0.21 (0.71)
700–1000 0.03 ±0.26 (0.40) 0.02 ±0.22 (0.44) 0.03 ±0.26 (0.42) 0.02 ±0.28 (0.56) 0.02 ±0.27 (0.45)
1000–1400 0.23 ±0.19 (0.01) 0.17 ±0.17 (0.02) 0.23 ±0.19 (0.01) 0.25 ±0.20 (0.01) 0.24 ±0.20 (0.01)
1400–1800 0.01 ±0.16 (0.55) 0.02 ±0.14 (0.38) 0.00 ±0.17 (0.53) 0.01 ±0.17 (0.55) 0.01 ±0.17 (0.58)
1800–1900 0.03 ±0.38 (0.45) 0.40 ±0.28 (1.00) 0.05 ±0.39 (0.40) 0.23 ±0.38 (0.89) 0.09 ±0.39 (0.69)
1860–1900 0.41 ±0.54 (0.94) 0.69 ±0.32 (1.00) 0.40 ±0.56 (0.93) 0.60 ±0.48 (0.99) 0.50 ±0.51 (0.98)
1900–2000 1.38 ±0.15 (1.00) 1.35 ±0.13 (1.00) 1.37 ±0.15 (1.00) 1.40 ±0.14 (1.00) 1.39 ±0.14 (1.00)
Amplitude ±8(7–11) ±6(4–8) ±9(7–11) ±8(6–10) ±8(6–11)
(0–1900; cm)
Errors are ±2. Parenthetical numbers indicate probability greater than 0–1700 CE average rate. Amplitude row indicates the median (5th–95th percentile) estimate of
the amplitude of variability over 0–1900 CE.
Dataset S1f. Rates of GSL change employing dierent data subsets (mm/yr; prior ML2,1)
Subset 0–700 700–1000 1000–1400 1400–1600 1600–1800 1800–1900 1900–2000
+All+GSL 0.10 ±0.10** 0.03 ±0.26 0.23 ±0.19*** 0.29 ±0.35** 0.28 ±0.31** 0.03 ±0.38 1.38 ±0.15***
+All-GSL 0.10 ±0.10** 0.04 ±0.26 0.23 ±0.19*** 0.29 ±0.35** 0.25 ±0.32*0.04 ±0.40 1.31 ±0.33***
+NWAtlantic+GSL 0.00 ±0.13 0.00 ±0.30 0.13 ±0.210.37 ±0.38** 0.21 ±0.380.07 ±0.53 1.37 ±0.16***
+NWAtlantic-GSL 0.00 ±0.13 0.00 ±0.30 0.12 ±0.210.37 ±0.38** 0.20 ±0.390.18 ±0.581.48 ±0.52***
-NWAtlantic+GSL 0.19 ±0.14*** 0.04 ±0.37 0.29 ±0.29** 0.02 ±0.54 0.34 ±0.44*0.04 ±0.47 1.37 ±0.16***
-NWAtlantic-GSL 0.18 ±0.14*** 0.04 ±0.37 0.29 ±0.29** 0.03 ±0.54 0.26 ±0.450.01 ±0.49 1.01 ±0.40***
+NEAtlantic+GSL 0.16 ±0.17** 0.01 ±0.45 0.16 ±0.350.04 ±0.65 0.48 ±0.53** 0.00 ±0.54 1.35 ±0.16***
+NEAtlantic-GSL 0.15 ±0.17** 0.01 ±0.45 0.17 ±0.350.04 ±0.65 0.28 ±0.540.08 ±0.57 0.51 ±0.54**
-NEAtlantic+GSL 0.06 ±0.110.04 ±0.27 0.18 ±0.20** 0.31 ±0.35** 0.19 ±0.350.17 ±0.481.39 ±0.16***
-NEAtlantic-GSL 0.06 ±0.110.04 ±0.27 0.18 ±0.20** 0.31 ±0.35** 0.20 ±0.360.28 ±0.511.56 ±0.40***
+NAtlantic+GSL 0.06 ±0.120.01 ±0.28 0.19 ±0.20** 0.33 ±0.37** 0.30 ±0.34** 0.04 ±0.41 1.36 ±0.15***
+NAtlantic-GSL 0.06 ±0.120.01 ±0.28 0.20 ±0.20** 0.33 ±0.37** 0.23 ±0.34*0.05 ±0.43 1.18 ±0.40***
-NAtlantic+GSL 0.12 ±0.15*0.04 ±0.41 0.21 ±0.31*0.02 ±0.56 0.16 ±0.530.29 ±0.701.41 ±0.17***
-NAtlantic-GSL 0.12 ±0.15*0.04 ±0.41 0.21 ±0.31*0.02 ±0.56 0.14 ±0.540.30 ±0.751.30 ±0.55***
+SAmerica+GSL 0.06 ±0.200.07 ±0.53 0.01 ±0.40 0.01 ±0.70 0.34 ±0.710.31 ±0.991.40 ±0.18***
+SAmerica-GSL 0.04 ±0.200.07 ±0.53 0.01 ±0.40 0.01 ±0.70 0.09 ±0.75 0.38 ±1.070.67 ±1.00*
-SAmerica+GSL 0.09 ±0.11** 0.02 ±0.26 0.23 ±0.19*** 0.30 ±0.35** 0.27 ±0.31** 0.01 ±0.38 1.37 ±0.15***
-SAmerica-GSL 0.09 ±0.11** 0.02 ±0.26 0.23 ±0.19*** 0.30 ±0.35** 0.24 ±0.32*0.01 ±0.40 1.26 ±0.34***
+AtlanticMediterranean+GSL 0.10 ±0.11** 0.00 ±0.27 0.33 ±0.21*** 0.45 ±0.37*** 0.26 ±0.34*0.01 ±0.42 1.35 ±0.15***
+AtlanticMediterranean-GSL 0.10 ±0.11** 0.00 ±0.27 0.34 ±0.21*** 0.45 ±0.37*** 0.19 ±0.340.04 ±0.44 1.11 ±0.39***
-AtlanticMediterranean+GSL 0.04 ±0.170.08 ±0.430.10 ±0.290.16 ±0.540.27 ±0.530.24 ±0.671.41 ±0.17***
-AtlanticMediterranean-GSL 0.04 ±0.170.07 ±0.430.10 ±0.290.16 ±0.540.27 ±0.530.38 ±0.701.58 ±0.54***
/*/**/*** = 67%/90%/95%/99% probability of sign. ’+’ indicates a data set consisting only of the sites in the named region, ’-’ a data set consisting of sites except
those in the named region. ’+GSL’ and ’-GSL’ represent the inclusion or exclusion of the Kalman smoother GMSL curve of ref. 12.
Kopp et al. Dataset 1 3
ii“CEGSL4b” — 2016/1/20 — 10:29 — page 4 — #21 i
i
iii
i
Dataset S1g. Rates of RSL change (mm/yr; prior ML2,1)
Site 0–1700 0–700 700–1400 1400–1800 1800–1900 1900–2000
Brittany-Arun 0.85 ±0.14*** 0.99 ±0.25*0.65 ±0.25*0.93 ±0.350.75 ±0.52 1.65 ±0.49***
Brittany-Tresseny 0.59 ±0.47*** 0.73 ±0.51*0.39 ±0.52** 0.68 ±0.570.50 ±0.65 1.39 ±0.61***
Christmas Island 0.06 ±0.07** 0.01 ±0.190.10 ±0.21 0.15 ±0.320.04 ±0.94 1.19 ±0.74***
Connecticut-Barn Island 0.91 ±0.22 1.06 ±0.20 1.09 ±0.62 2.69 ±0.45
Connecticut-East River Marsh 0.97 ±0.07*** 1.02 ±0.150.91 ±0.141.04 ±0.211.09 ±0.61 2.73 ±0.47***
Connecticut-Indian River 0.91 ±0.24*** 0.96 ±0.280.84 ±0.270.97 ±0.331.08 ±0.672.71 ±0.52***
Cook Islands-Aitutaki 0.58 ±0.24 0.54 ±0.37 1.10 ±0.96 1.25 ±0.87
Cook Islands-Rarotonga 0.53 ±0.36 1.20 ±0.90 1.23 ±0.79
Denmark-Ho Bugt 0.52 ±0.12*** 0.67 ±0.24*0.32 ±0.23** 0.60 ±0.370.71 ±0.89 1.43 ±0.78***
Florida-Nassau 0.40 ±0.06*** 0.43 ±0.150.38 ±0.15 0.39 ±0.27 0.49 ±0.83 1.82 ±0.49***
Greenland-Aasiaat 0.98 ±0.33 0.46 ±1.02 2.02 ±1.05
Greenland-Nanortalik 1.31 ±1.05 2.67 ±1.09
Greenland-Sisimiut 0.66 ±0.36 0.10 ±1.03 1.67 ±1.05
Iceland-Vioarholmi 0.65 ±0.08*** 0.82 ±0.20** 0.47 ±0.22*0.66 ±0.35 0.89 ±0.891.96 ±0.77***
Isle of Wight-Newtown Estuary 0.59 ±0.33*** 0.74 ±0.38*0.38 ±0.38** 0.69 ±0.490.56 ±0.67 1.43 ±0.61***
Israel-Casesarea 0.01 ±0.08 0.14 ±0.16** 0.22 ±0.20** 0.01 ±0.39 0.05 ±1.07 1.43 ±1.03***
Louisiana-Lydia 0.60 ±0.18 0.49 ±0.30 0.25 ±0.96 3.41 ±0.83
Massachusetts-Barnstable 1.16 ±0.26*** 1.22 ±0.311.07 ±0.291.24 ±0.361.23 ±0.75 2.81 ±0.57***
Massachusetts-Revere 0.62 ±0.08*** 0.69 ±0.170.53 ±0.150.71 ±0.250.69 ±0.69 2.35 ±0.45***
Massachusetts-Wood Island 0.54 ±0.09*** 0.61 ±0.190.45 ±0.130.63 ±0.240.64 ±0.69 2.23 ±0.47***
New Jersey-Cape May Courthouse 1.54 ±0.10*** 1.64 ±0.16*1.49 ±0.171.49 ±0.241.78 ±0.623.87 ±0.43***
New Jersey-Cheesequake Marsh 1.46 ±0.33*** 1.53 ±0.35*1.38 ±0.361.47 ±0.42 1.72 ±0.723.47 ±0.50***
New Jersey-Leeds Point 1.53 ±0.06*** 1.63 ±0.13** 1.45 ±0.131.52 ±0.24 1.77 ±0.633.76 ±0.41***
New Zealand-Blueskin Bay 0.03 ±0.21 0.10 ±0.290.18 ±0.300.07 ±0.40 0.14 ±0.90 1.47 ±0.56***
New Zealand-Pounawea 0.20 ±0.33 0.30 ±0.40 0.56 ±0.84 2.08 ±0.60
North Carolina-Croatan National Forest 0.66 ±0.30 0.45 ±0.34 0.83 ±0.67 2.77 ±0.57
North Carolina-Hatteras Island 0.99 ±0.60 1.21 ±0.79 3.40 ±0.76
North Carolina-Sand Point 1.12 ±0.04*** 1.08 ±0.101.24 ±0.09*** 0.99 ±0.16*1.08 ±0.50 3.33 ±0.47***
North Carolina-Tump Point 1.06 ±0.08*** 1.05 ±0.16 1.16 ±0.15*0.91 ±0.14** 1.33 ±0.483.47 ±0.41***
North Carolina-Wilmington 0.68 ±0.17*** 0.69 ±0.23 0.73 ±0.220.58 ±0.300.90 ±0.732.58 ±0.54***
Nova Scotia-Chezzetcook 1.78 ±0.13*** 1.87 ±0.241.65 ±0.231.89 ±0.221.66 ±0.53 3.14 ±0.45***
Rio de Janeiro-Arraial do Cabro 0.79 ±0.28*** 0.68 ±0.350.94 ±0.350.80 ±0.47 0.90 ±1.08 0.90 ±0.95***
Rio de Janeiro-Buzios 0.66 ±0.25*** 0.54 ±0.330.80 ±0.330.66 ±0.45 0.76 ±1.07 1.03 ±0.94***
Rio de Janeiro-Frade 0.63 ±0.30*** 0.51 ±0.370.77 ±0.370.64 ±0.48 0.73 ±1.07 1.08 ±0.94***
Rio de Janeiro-Ilha Grande 0.87 ±0.47*** 0.75 ±0.531.02 ±0.510.89 ±0.60 1.03 ±1.14 0.86 ±1.00***
Rio de Janeiro-Itaipu-Acu 0.20 ±0.16*** 0.09 ±0.270.35 ±0.270.21 ±0.41 0.31 ±1.04 1.53 ±0.88***
Rio de Janeiro-Mangaratiba 0.64 ±0.24*** 0.52 ±0.320.79 ±0.320.66 ±0.45 0.79 ±1.06 1.09 ±0.91***
Rio de Janeiro-Parati-Mirim 0.85 ±0.42*** 0.73 ±0.471.00 ±0.480.87 ±0.57 1.02 ±1.12 0.87 ±0.99***
Rio de Janeiro-Tarituba 0.91 ±0.43*** 0.79 ±0.481.06 ±0.490.93 ±0.57 1.08 ±1.13 0.81 ±1.00***
Santa Catarina-Cape of Santa Marta 0.58 ±0.24*** 0.46 ±0.320.74 ±0.320.59 ±0.45 0.81 ±1.10 1.00 ±1.01***
Santa Catarina-Ponta de Itapiruba 0.49 ±0.54** 0.37 ±0.580.65 ±0.580.51 ±0.66 0.73 ±1.20 1.10 ±1.10***
Scotland-Kyle of Tongue 0.36 ±0.27 0.16 ±0.38 0.25 ±0.85 0.83 ±0.71
Scotland-Loch Laxford 0.07 ±0.25 0.13 ±0.36 0.03 ±0.85 1.13 ±0.71
Scotland-Wick 0.15 ±0.12*** 0.01 ±0.24** 0.35 ±0.23** 0.14 ±0.36 0.22 ±0.83 0.85 ±0.67***
Seychelles-Barbarons 0.44 ±0.15*** 0.54 ±0.260.30 ±0.260.44 ±0.41 0.41 ±1.11 1.81 ±1.05***
South Africa-Groenvlei 0.03 ±0.36 0.36 ±0.47 0.10 ±1.04 1.48 ±0.85
South Africa-Kariega Estuary 0.42 ±0.28 0.79 ±0.37 0.49 ±1.01 1.76 ±0.89
South Africa-Langebaan 0.02 ±0.16 0.12 ±0.27*0.25 ±0.25** 0.08 ±0.400.11 ±1.05 1.31 ±0.89***
South West England-Thurlestone 0.95 ±0.16*** 1.10 ±0.29*0.73 ±0.23** 1.04 ±0.370.93 ±0.70 1.78 ±0.60***
Spain-Muskiz Estuary 0.92 ±0.51*** 1.02 ±0.550.75 ±0.55*1.00 ±0.630.90 ±0.98 1.87 ±0.78***
Spain-Urdaibai Estuary 0.52 ±0.41*** 0.61 ±0.440.34 ±0.44*0.60 ±0.570.51 ±0.82 1.45 ±0.70***
Tasmania-Little Swanport 0.98 ±1.19 1.88 ±0.76
/*/**/*** = 67%/90%/95%/99% probability of sign (for 0–1700 CE) or of being distinct from the 0–1700 CE rate. Rates not shown when there are no observations
within 0.5 degrees of a site. Probabilities not shown where no observations within 0.5 degrees of a site predate 0 CE.
Kopp et al. Dataset 1 4
ii“CEGSL4b” — 2016/1/20 — 10:29 — page 5 — #22 i
i
iii
i
Dataset S1h. Probability observed GSL rise since 1900 CE exceeded counterfactual projection
Scenario 1 Scenario 2
calibrated against: calibrated against:
Yea r Mann et al. (ref. 1) Marcott et al. (ref. 2) Mann et al. (ref. 1) Marcott et al. (ref. 2)
1910 0.77 0.37 0.02 0.02
1920 0.94 0.51 0.13 0.11
1930 1.00 0.72 0.60 0.54
1940 1.00 0.88 0.96 0.94
1950 1.00 0.95 1.00 1.00
1960 1.00 0.98 1.00 1.00
1970 1.00 0.99 1.00 1.00
1980 1.00 0.99 1.00 1.00
1990 1.00 1.00 1.00 1.00
Dataset S1i. Semi-empirical projections of 21st century sea-level rise with dierent calibrations (cm)
50 17–83 5–95 50 17–83 5–95 50 17–83 5–95
Summary Mann et al. (ref. 1) Marcott et al. (ref. 2)
RCP 2.6 38 28–51 24–61 38 29–50 25–59 38 28–51 24–61
RCP 4.5 51 39–69 33–85 51 39–66 34–81 52 39–69 33–85
RCP 8.5 76 59–105 51–131 75 59–99 52–121 78 60–105 52–131
Values with respect to year 2000 baseline. Results across the three temp erature calibration sets show median of medians, minimum of 5th percentiles, and maximum of
95th percentiles.
Dataset S1j. Prior and posterior distributions P( )for the parameters
Parameter Prior Mann et al. (ref. 1) Marcott et al. (ref. 2)
a U(0,20) mm/yr/K 4.0(3.2,5.4) 4.7(3.4,7.0)
c(500 CE) U(10,10) mm/yr 0.22 (0.10,0.42) 0.05 (0.02,0.08)
c(2000 CE) U(2,2) mm/yr 0.14 (0.05,0.29) 0.03 (0.01,0.06)
T0(2000 CE) <T (2000 to 1800 CE)>+U(0.6,0.6) 0.25 (0.19,0.89) 0.03 (0.42,0.60)
T0(500 CE) N(<T (500 700 CE)>, (0.2K)
2) 0.17 (0.11,0.23) 0.09 (0.17,0.01)
T0(2000 CE) 0.05 (0.12,0.07) 0.04 (0.10,0.16)
log U(30,300) yrs 174 (87,366) 102 (64,203)
clog U(1000,20000) yrs 4175 (1140,17670) 3392 (1124,16155)
N(µ, 2)denotes a normal distribution around µwith the standard deviation .U(x1,x
2)is a uniform distribution between
x1and x2. Ranges shown for posteriors are 5th–95th percentiles. Temperatures are relative to the 1850–2000 CE average.
Kopp et al. Dataset 1 5
... To compute the climatic effects on LOD, we use the available data set of Kopp et al. (2016). These data represent the sea-level variation due to barystatic processes (Gregory et al., 2019) as well as their uncertainties. ...
... spacecenter.dk/files/magnetic-models/COV-OBSx2/. The climatic data to compute LOD from barystatic processes are freely available at Kopp et al. (2016). The LOD data analyzed in the study can be accessed at Kiani . ...
Article
Full-text available
Length of Day (LOD) observations in the range 720 BCE to 2020—derived from lunar occultation and eclipse records—feature a secular trend and various long‐period fluctuations. While recent estimates show that the secular trend is caused by the combination of lunar tidal friction and glacial isostatic adjustment, the causes of long‐period fluctuations remain ambiguous. We first compute the climatic effects and show that they are anti‐correlated with the observed fluctuations and their amplitude is ∼ {\sim} 10 smaller. Then, we focus on core dynamics and solve for simplified equations of magnetohydrodynamics, namely tangential geostrophy, using Bayesian Physics‐Informed Neural Networks (BPINNs) and independent archeomagnetic and modern geomagnetic observations. Within the observation and reconstruction uncertainty we can reconcile the LOD observations with reconstructions of BPINNs. Furthermore, we demonstrate that LOD variations reconstructed by dynamics of Magneto‐Archimedes‐Coriolis waves do not explain the observed fluctuations. These results have considerable implications for internal and external geodynamics.
... Sea-level change has captivated widespread attention as anthropogenic warming has resulted in an acceleration of global rise from near zero in most of the Common Era (Kopp et al., 2016) to 1-2 mm/yr in the 20th century (Barnett, 1990; best estimate 1.4 ± 0.3 mm/yr; Hay et al., 2015), to 3.5 ± 0.4 mm/yr in 2024 (Hamlington et al., 2023;Nerem et al., 2010;Sea Level Research Group, 2025). The stratigraphic record also speaks of ancient worlds of dramatically different sea levels, with vast areas of the continents inundated on an ice-free planet as recently as the Cretaceous to Eocene Greenhouse worlds and with large areas of the continental shelf exposed when sea level was lowered by ~130 m at the last glacial maximum ca. ...
... These areas are especially vulnerable to flooding and erosion from major storms. As the global sea level continues to rise [3][4][5], coastal communities face an increasing risk of more frequent and destructive inundation and erosion due to storm events [6,7]. ...
Article
Full-text available
Sandy barrier systems are highly dynamic, with the most significant natural morphological changes to these systems occurring during high-energy storm conditions. These systems provide a range of economic and ecosystem benefits and protect inland areas from flooding and storm impacts, but the persistence of many coastal barriers is threatened by storms and sea-level rise (SLR). This study employed observations and modeling to examine recent and potential future influences of storms on a sandy coastal barrier system in Nauset Beach, MA. Drone-derived imagery and digital elevation models (DEMs) of the study area collected throughout the 2023–2024 winter revealed significant alongshore variability in the geomorphic response to storms. Severe, highly localized erosion (i.e., an erosional “hotspot”) occurred immediately south of the Nauset Bay spit as the result of a group of storms in December and January. Modeling results demonstrated that the location of the hotspot was largely controlled by the location of a break in a nearshore sandbar system, which induced larger waves and stronger currents that affected the foreshore, backshore and dune. Additionally, model simulations of the December and January storms assuming 0.3 m (1 ft) of SLR showed the system to be relatively resistant to major geomorphic changes in response to an isolated storm event, but more susceptible to significant overwash and breaching in response to consecutive storms. This research suggests that both very strong isolated storm events and sequential moderate storms pose an enhanced risk of major overwash, breaching, and possibly inlet formation today and into the future, raising concern for adjacent communities and resource managers.
... In this sense, floods associated with sea level rise represent a strong threat to the exposed coastal population, housing and service infrastructure, related economic activities, coastal ecosystems, and the loss of their environmental services (Kopp et al., 2016;Adeel et al., 2020;. Therefore, the assessment of potential flood-related damage is a key requirement in coastal zone planning and management (Bates et al., 2005;Hallegatte et al., 2011;Gallient et al., 2014;Seenath, 2015;Seenath et al., 2016;Gallient, 2016;Rey et al., 2020), and takes special relevance to establish plans for urban development, disaster prevention, emergency response and management and recovery strategies, especially in areas with strong environmental and economic importance, such as Campeche. ...
Article
The city of San Francisco de Campeche, located in the Yucatan Peninsula, Mexico; is vulnerable to coastal flooding due to its geographical location and low altitude terrain. Under these inherent site conditions, sea level rise associated with climate change represents a potential threat to people's property and goods in the coming decades. In this work, three scenarios of sea level rise in the coastal area of the city of San Francisco de Campeche were evaluated through numerical simulation and using wave and wind data obtained from the ERA5 reanalysis model as inputs: astronomical tide data and a high-resolution topobathymetric model. The scenarios evaluated correspond to the periods 2031-2050, 2046-2065, and 2081-2100 reported in the latest IPCC assessment report. Damage to people's property and goods was analyzed using the CENAPRED methodology based on the type of housing. The results allow identifying that the area of the old neighborhood of San Román, Colonia Miramar, Pedro Sainz de Baranda, Adolfo Ruiz Cortinez, Resurgimiento avenues, and the Campeche-Merida and Campeche-Champoton coastal highway, as well as the federal and state government offices they will be the most affected areas by sea level rise with economic damages exceeding 13,860,322Mexicanpesos(13,860,322 Mexican pesos (USD 805,832.67) under the 2031-2050 scenario, 14,706,754Mexicanpesos(14,706,754 Mexican pesos (USD 855,043.83) in the 2046-2065 scenario and 22,536,250Mexicanpesos(22,536,250 Mexican pesos (USD 1,310,247.09) in the 2081-2100 scenario; and the Los Petenes Biosphere Reserve as one of the ecological zones with the greatest extension of flooding; as well as downtown areas of the city that currently have residential, commercial, recreational, and port uses. Considering the three scenarios and effects on the population and housing, the 2081-2100 scenario is the one that generates the greatest flooding and with it a greater economic loss that exceeds 12 million dollars compared to the 2046-2050 scenario
... The lack of sediment sources, SLR, sand mining and destruction of natural defence lines due to increasing human occupation are projected to create additional pressure on these areas and accelerate erosion mechanisms (Mangor et al., 2017). It is estimated that the global mean sea level has already increased by 13-20 cm since pre-industrial times (Kopp et al., 2016), accelerating since the 1990s (Watson et al., 2015), which has already contributed to coastal recession in some areas of Europe (EUROSION, 2003;Leatherman et al., 2000;Mentaschi et al., 2018), and increased the susceptibility to coastal hazards. ...
Technical Report
Full-text available
Some of the most disruptive effects of climate change are projected to be felt along the coastlines. From flooding to extreme coastal erosion, future changes in coastal dynamics are particularly feared, especially if combined with sea level rise, tides, storm surges and changes in wave climate. Coastal areas are amongst the most vulnerable regions to climate change, comprising important populational centres and economically relevant hubs. The portion of total population living in coastal areas has rapidly increased in the last decades, being estimated that at least 10% of the current world’s population lives near the coast, less than 10 m above sea-level. In Portugal, data from the CENSOS2011 shows that 14% of the national population lives within 2 km of the sea, with the most recent update (CENSOS2021) pointing to increases in the Lisbon and Algarve regions, of 1.7% and 3.7%, respectively, in comparison with 2011. Rising sea levels, together with the effects of tides, storm surges and extreme waves are considered key-drivers of coastal hazards, threatening coastal infrastructures, ecosystems, and communities. The increase in human pressure along the Portuguese coastlines calls for a reliable, long-term coastal vulnerability assessment, paramount for effective coastal management, sustainable development, adaptation, and impact mitigation strategies. In the context of an increasing need for accurate physical and socioeconomic coastal vulnerability assessments, and incorporated in the National Roadmap for Adaptation XXI, we present a thorough and comprehensive assessment of future projected hydro-morpho-dynamical changes along the Portuguese coastlines. Future shoreline evolution and extreme coastal flooding projections are obtained, through high-resolution hydro- and morpho-dynamic modelling, for five coastal key-locations, selected due to their higher currently perceived vulnerability to climate change (based on historical records). Ensemble-based projections forced by Coupled Model Intercomparison Project phase 5 (CMIP5) Global Climate Models (GCMs), are used to drive an innovative methodology, focused on dealing with the multivariate challenges of an accurate coastal vulnerability assessment for Portugal, aiming to accurately assess the extension of future projected extreme coastal flooding. Two Representative Concentration Pathway scenarios are considered, namely the RCP4.5 and the RCP8.5. These baseline results are used to train a parametric approach designed for the complete, national-scale coastal vulnerability assessment, supported by a composed coastal vulnerability index. At a local scale, our results indicate that future nearshore wave action, projected to become more northerly and less energetic, is expected to lead to northward beach rotations especially along the northern and central Portuguese coastal stretches. Nevertheless, the impact of SLR is shown to lead to consistent shoreline retreats throughout all analyzed key-locations. Such results are in agreement with several studies 21 indicating that while wave action is projected to dominate morphological response until the mid-21st century, SLR should become the main driver of shoreline evolution beyond that time-frame. Final projected shoreline retreats are shown to locally reach 100 m (120 m) by 2100 under RCP4.5 (RCP8.5) at Ofir, 200 m (210 m) at Costa Nova, 140 m (150 m) at Cova Gala, 290 m (300 m) along Costa da Caparica, and 65 m (80 m) in Praia de Faro. The projected lost areas between the reference (2018) and future mean shorelines range between 0.088 km2 and 0.184 km2 (0.118 km2 and 0.197 km2) by 2100, under RCP4.5 (RCP8.5), the smallest (greatest) losses expected to take place at Faro and Cova Gala (Costa Nova). Throughout all key-locations (approximately 14 km of coastline), the cumulative amount of projected lost area from 2018 to 2100 ascends to 0.786 km2 (2100 under RCP8.5), relevant when compared to the historical nationwide area lost to the sea between 1958 and 2021, which amounted to 13.5 km2 for over 980 km of coastline. The synchronized action of extreme total water levels, resulting essentially from SLR, but also from the joint occurrence of high spring tides or storm surge conditions, in the context of weaker natural protection structures due to erosion, is shown to lead to unprecedented coastal flooding in the future. Throughout the five key-locations, the future projected threatened area, expected to become flooded under extreme conditions, is projected to ascend to 0.657 km2 (0.738 km2) by 2070 under RCP4.5 (RCP8.5), and 0.841 km2 (1.47 km2) by 2100 under RCP4.5 (RCP8.5). Based on the dynamical modelling at the five key-locations, a parametric approach is calibrated to characterize coastal retreat, flooding and the overall vulnerability along the entire Portuguese coastline. The coastal vulnerability index, divided into three levels (low, moderate and high), is inversely related to the projected flooding extent, so that areas under high CVI are the ones showing increased vulnerability to less extreme (more frequent) events, and vice-versa. Finally, the ocean-facing areas under CVI along Mainland Portugal are projected to ascend to 41.7 km2 (2070 under RCP4.5), 49.7 km2 (2070 under RCP8.5), 54.7 km2 (2100 under RCP4.5) and 55.9 km2 (2100 under RCP8.5). These areas, related to episodically flooded territory, are projected to amount to 3.09, 3.68, 4.05 and 4.14 times the area observed to have been lost between 1958 and 2021 (13.5 km2). However, when considering inland waters, an additional value between 514 km2 and 548 km2 (2070 under RCP4.5 and 2100 under RCP8.5, respectively) must be considered. Therefore, for all types of coastlines along Mainland Portugal, the future area under CVI is projected to ascend to 604 km2 by 2100, under the RCP8.5 scenario.
... 143, Louisiana, 2017]. Note that the flood probability in 2065 is almost certainly larger than the average annual exceedance probability due to temperature-driven increases in sea level and storm surge intensities [Grinsted et al., 2013;Kopp et al., 2016]. Note further that the failure probabilities calculated here consider only the protection offered by the levee system; other adaptive and protective measures may increase the effective return period. ...
Preprint
Future sea-level rise drives severe risks for many coastal communities. Strategies to manage these risks hinge on a sound characterization of the uncertainties. For example, recent studies suggest that large fractions of the Antarctic ice sheet (AIS) may rapidly disintegrate in response to rising global temperatures, leading to potentially several meters of sea-level rise during the next few centuries. It is deeply uncertain, for example, whether such an AIS disintegration will be triggered, how much this would increase sea-level rise, whether extreme storm surges intensify in a warming climate, or which emissions pathway future societies will choose. Here, we assess the impacts of these deep uncertainties on projected flooding probabilities for a levee ring in New Orleans, Louisiana. We use 18 scenarios, presenting probabilistic projections within each one, to sample key deeply uncertain future projections of sea-level rise, radiative forcing pathways, storm surge characterization, and contributions from rapid AIS mass loss. The implications of these deep uncertainties for projected flood risk are thus characterized by a set of 18 probability distribution functions. We use a global sensitivity analysis to assess which mechanisms contribute to uncertainty in projected flood risk over the course of a 50-year design life. In line with previous work, we find that the uncertain storm surge drives the most substantial risk, followed by general AIS dynamics, in our simple model for future flood risk for New Orleans.
Article
Full-text available
Conventional computational models of climate adaptation frameworks inadequately consider decision-makers’ capacity to learn, update, and improve decisions. Here, we investigate the potential of reinforcement learning (RL), a machine learning technique that efficaciously acquires knowledge from the environment and systematically optimizes dynamic decisions, in modeling and informing adaptive climate decision-making. We consider coastal flood risk mitigations for Manhattan, New York City, USA (NYC), illustrating the benefit of continuously incorporating observations of sea-level rise into systematic designs of adaptive strategies. We find that when designing adaptive seawalls to protect NYC, the RL-derived strategy significantly reduces the expected net cost by 6 to 36% under the moderate emissions scenario SSP2-4.5 (9 to 77% under the high emissions scenario SSP5-8.5), compared to conventional methods. When considering multiple adaptive policies, including accomodation and retreat as well as protection, the RL approach leads to a further 5% (15%) cost reduction, showing RL’s flexibility in coordinatively addressing complex policy design problems. RL also outperforms conventional methods in controlling tail risk (i.e., low probability, high impact outcomes) and in avoiding losses induced by misinformation about the climate state (e.g., deep uncertainty), demonstrating the importance of systematic learning and updating in addressing extremes and uncertainties related to climate adaptation.
Book
Full-text available
A comprehensive and self-contained introduction to Gaussian processes, which provide a principled, practical, probabilistic approach to learning in kernel machines. Gaussian processes (GPs) provide a principled, practical, probabilistic approach to learning in kernel machines. GPs have received increased attention in the machine-learning community over the past decade, and this book provides a long-needed systematic and unified treatment of theoretical and practical aspects of GPs in machine learning. The treatment is comprehensive and self-contained, targeted at researchers and students in machine learning and applied statistics. The book deals with the supervised-learning problem for both regression and classification, and includes detailed algorithms. A wide variety of covariance (kernel) functions are presented and their properties discussed. Model selection is discussed both from a Bayesian and a classical perspective. Many connections to other well-known techniques from machine learning and statistics are discussed, including support-vector machines, neural networks, splines, regularization networks, relevance vector machines and others. Theoretical issues including learning curves and the PAC-Bayesian framework are treated, and several approximation methods for learning with large datasets are discussed. The book contains illustrative examples and exercises, and code and datasets are available on the Web. Appendixes provide mathematical background and a discussion of Gaussian Markov processes.
Article
Full-text available
No important instrumental modifications have been made since publication of our last list (Yale VII). As before, we are indebted to many collaborators for samples and for assistance with descriptions and comments. Technical assistance has been provided by George Young, Carolyn Haupt, and Jonathan Wood. The generous support of the National Science Foundation, under grants G-19080 (to Stuiver), G-19335 (to Deevey), and G-24049 (to Rouse) is gratefully acknowledged.
Article
Full-text available
The following list includes most of the measurements made since publication of Yale V; some measurements, such as a series collected in Greenland by A. L. Washburn, are withheld pending additional information or field work that will make better interpretations possible. In addition to radiocarbon dates of geologic and/or archaeologic interest, we give in the third part of the paper, for the first time since 1954 (Deevey and others, 1954), recent assays of C 14 in lake waters and other lacustrine materials, now normalized for C 13 content. Some of these C 13 values have been published separately (Oana and Deevey, 1960), but most have not.
Article
The radiocarbon dating laboratory of Queens College, CUNY, was established with the joint support of the Research Foundation of the City University of New York and the departments of Anthropology, Biology, Chemistry, and Earth and Environmental Sciences of Queens College. The laboratory has been operational since June, 1975 and is located in the basement of a two-story concrete and brick building.
Article
The following list includes radiocarbon analyses of samples related to studies of Holocene sea levels completed since the publication of the last list (R, 1980, v 22, p 1073–1083). Sample preparation and counting for liquid scintillation samples remain the same. However, an additional gas-proportional facility was added in 1981 to handle the analyses of small samples, some of which are included in this list. The new system consists of two 660cc OFHC copper counters built at Queens College. Samples are counted over at least two 2800 minute intervals alternating with backgrounds and standards counted over 1400 minute intervals. Ages are based on the Libby half-life of 5568 years and include 1 σ standard deviations of sample, standard, and background activities.
Article
Dates obtained since our last date list (Yale III∗), and up to the end of 1958, are given here. All are based on duplicate measurements of CO 2 in a proportional counter. Grateful acknowledgment is made to the National Science Foundation, which has supported our work through grant no. G-2278. Participation of Väinö Hoffren during 1956–1957 was made possible through an arrangement with the Geological Survey of Finland. The technical assistance of Julius Kovats has been invaluable.
Conference Paper
We give a basic introduction to Gaussian Process regression models. We focus on understanding the role of the stochastic process and how it is used to define a distribution over functions. We present the simple equations for incorporating training data and examine how to learn the hyperparameters using the marginal likelihood. We explain the practical advantages of Gaussian Process and end with conclusions and a look at the current trends in GP work.