Vous êtes sur la page 1sur 8

underestimated in many models.

Concluding, there is
a strong need for MME and PPE studies into
bistability of the AMOC, accounting for freshwater
influx from rivers and melting of the Greenland ice
sheet. To date, these intercomparisons have proven
too computationally expensive, but EMICs can provide
an accurate alternative (Stouffer et al. 2006, Gregory
et al. 2005). Stommels salt-advection feedback is an
important link which is hard to trace in the more
comprehensive AOGCMs and their ensemble
simulations, because of their dedication to presentclimate stability and overall underappreciation of
high-integration processes of feedback, which EMICs
do take into account (Hofmann and Rahmstorf 2009,
Claussen et al. 2002).

external basins. The conceptual setup is shown in


Figure 8.
The elegance of Stommels model is its
commitment to symmetry. The temperatures and
salinity of box one are chosen as the exact negatives
of the other box, with zero the benchmark
temperature.

4. Stommels salt-advective box model


The discussed developments in climate research
seem to suggest a need for widening the spectrum of
model complexity, both in order to assess the
integration of physical processes such as feedbacks,
as well as in order to explore a wider parameter space
in ensemble simulations. Particularly the role of salt
advection in the Meridional Overturning Circulation
causes a continuous attention to simple models that
are conceptually strong in reproducing the
characteristic freshwater hysteresis response that is
found in some EMICs and AOGCMs. Simple climate
models of the ocean often involve the representation
of ocean basins in the form of interconnected boxes
through which water circulates due to density
differences linearly related to the waters
temperature and salinity. Seminal work in this respect
is Stommels two-box model of thermohaline
circulation (1961).
Because in the real ocean salinity and
temperature can have entirely different flux
mechanisms at different spatial dimensions, a clear
demarcation of ocean volumes can be effective in
analysing feedback mechanisms between them.
Stommel did this by representing the MOC with
thermohaline circulation through two boxes,
connected at the top and at the bottom. Each box is in
turn connected to an external basin of constant
reference density and temperature. One box could
represent the poles or the North Atlantic, where the
other could represent the tropics, depending on the
initial conditions, thermohaline transfer coefficients,
flow parameters and reference properties of the
10

Figure 8. Stommels two-box model setup

This gives two simple partial differential equations


that can be solved analytically:

Here, c and d represent the transfer coefficients


for temperature and salinity respectively. By making
the system dimensionless, Stommel shows that the
system depends on only three parameters in the
following equations:

Temperature and salinity are denoted by y and x,


represents a measure of resistance to flow between
the basins, the temperature-salinity relaxation time
(the ratio of their transfer coefficients) and R the
pycnal effects (the strength of density difference).
Solution of the setup yields the temperaturesalinity diagram shown in Figure 9. Clearly visible in
the diagram are the three equilibrium points, one of
which is the hysteresis bifurcation point, discussed
earlier. Perfect mixing within the basins is a crucial
elements. The imaginary mixing rods would have as
their counterpart in reality turbulent convective
mixing in the ocean. Renewed interest in the salt-

advection feedback mechanism has prompted


laboratory experiments in which thermohaline
hysteresis is reproduced (Whitehead 2009). Using an
altered setup of one freshwater reservoir and one
active basin, in which top saltwater influx and bottom
heating would cause fast convective mixing,
hysteresis is observed (Figure 10). Other authors
looking to conceptually re-examine Stommels box
model have pointed towards the prevalence of winddriven circulation, which could influence the position
of the bifurcation point (Guan and Huang 2008). For
the purpose of examining thermohaline hysteresis
effectively, we will not consider wind-stress effects in
the following model.

Figure 10. Thermohaline hysteresis reproduced in


a box experiment. Temperature and salinity are
plotted as a function of the bottom heating bath
temperature. One can clearly see bistability in the
system. Adopted from Whitehead (2008).

Four-box salt-advection model with entrainment


For the purpose of addressing the issues put
forward earlier, a more complex version of Stommels
model is presented and evaluated upon its core
results. There are a few changes made in the model
setup:
- Each basin is divided internally by a
thermocline, effectively giving a total of four
boxes
- The thermocline allows turbulent property
exchange through diffusive entrainment,
parametrised with constant flow w and v,
independent from the bottom and top flow q
- Each box is distinctly represented in the
model, resulting in four variables each for
salinity and temperature
- The main basins are connected to external
basins only above the thermocline

Figure 9. In this diagram of Stommels box of


thermohaline circulation, a few integral curves
show the three equilibrium points the system can
have, depending on the model parameters , R and
. a)is a temperature-dominated stable node, c) a
salinity-dominated spiral node and b) an unstable
saddle point, also known as the bifurcation point
beyond which the system can go from salinity
driven to temperature driven, or vice-versa.
Density anomaly lines are drawn to show the T-S
ratio.

The purpose of the model exercise is to study the


effects of complexification on Stommels core result,
and subsequently suggest modelling avenues that
modify the pioneer model to resemble the actual
physics and geometry of the MOC. This is dependent
on whether thermohaline hysteresis and bistability
are reproduced despite the new model geometry and
entrainment. In comparison to the original model,
initial conditions will be simulated that imitate
symmetry between the basins. The setup chosen is
seen in Figure 11.

The model used is a four-box temperature and


salinity construction derived from Stommels original
two-box model. It consists of 2 main basins of equal
height and surface area, where each basin has a
thermocline which vertically separates two internal
basins with different salinity and temperature, which
interact through entrainment.

11

4 4 A

1 1 A

q
T4 S4

T1 S1
V4

V1 = Ah1
v

w
T3 S3

T2 S2
V3

V2

Figure 11. Setup of four-box salt-advective model with separate variables and entrainment along a
simulated thermocline. Entrainment can work to either counteract or contribute to the overall flow q.

The flow q, determined by the density difference


of the two basins, must be equal from each box to the
next so that the volume of the basins remains the
same. In addition to this regular flow, the two basins
have internal entrainment flow at a specific height h
which is dependent on the temperature diffusion
coefficient and the surface of the thermocline. The
two top basins also simulate property exchange with
the atmosphere through external basins with a fixed
temperature and salinity. The equations governing
this system are derived from standard equations of
fluid property mixing (with V the volume of a basin):
=

(T

T )
dT
= (T T ) =
dt
V
dt

T1 V

The main flow q can be in two directions and in either


situation the equations look very different. This
problem is solved by converting them into signdependent components:
+ ||
2
||
=
2

1
V
1
=
V
1
=
V
1
=
V

(T T ) + (T T ) + ||(T T ) +

(S S ) + (S S ) + ||(S S )
(S S ) + (S S ) + ||(S S )
(S S ) + (S S ) + ||(S S ) +

dA
( S )
V

p =k
p =k
= h + h
= h + h

(k + k ) = ( h + h ) ( h + h )
(k + k ) = ( h + h ) ( h + h )
Now, we can reduce both bottom and top flow to a
single resistance parameter:

cA
( T )
V

(T T ) + (T T ) + ||(T T )

k=k +k :

(T T ) + (T T ) + ||(T T )
(T T ) + (T T ) + ||(T T ) +

dA
( S )
V

Solving for q, this gives:

The exchange of property at the surface of basins 1


and 4 is linearly dependent on the product of the
surface area A and the salinity and temperature
transfer coefficients d and c. The system of equations
then becomes:
=

(S S ) + (S S ) + ||(S S ) +

p
p
p p
p p

dT
dt
dT
dt
dT
dt
dT
dt

1
V
1
=
V
1
=
V
1
=
V

In order to determine the flow parameters, we


define the main flow from the hydrostatic balance and
the internal flow in the two basins through constants
of entrainment v and w. We calculate the pressure
balance that determines the bottom and top flow
using p1 and p4 as the basin pressures at the surface,
and p2 and p3 at the bottom. Surface flow is
constrained by parameter kt and bottom flow by kb.
The pressure difference determining the flow at the
bottom of the two basins is defined by p2-p1 and p3-p4
respectively:

dtqT2

(dtT + T V dtT )
V

dS
dt
dS
dt
dS
dt
dS
dt

cA
( T )
V

12

We also introduce a simple form of the equation of


state based on the temperature and salinity transfer
coefficients and :

Where the z parameter indicates the relative position


of the thermocline in each main basin. We also
determine the transfer coefficients using an abovethermocline temperature diffusion time tt:

= (1 T + S )

Hz
(1 + z )t
Hz
d=
(1 + z )t
c=

We then end up with a flow given by:


=

(T h + S h T h + S h )
(T h + S h T h + S h )

Definition in terms of dimensionless variables


In order to evaluate the model, we define the
previously introduced parameters k, i, i, d, c, and
such that they render the equations dimensionless.
The main interaction the model is concerned with is
the surface property exchange, because it represents
the systems external forcing. To reduce the models
complexity, we choose the external basin properties
such that they become 1 = 4 = , 1 = 4 = . We
can then make temperature and salinity
dimensionless by taking them relative to this
difference parameter:

Here, represents the temperature-salinity relaxation


time:

Az
( + R )
(1 + z )
A
( + R )
+
(1 + z )
A

( + R )
(1 + z )
Az

( + R )
(1 + z )

When we then introduce the parameter = ct, which


corresponds to the temperature dependent timescale,
the system of equations becomes:

In the choice of parameters later on, and are in the


same order of magnitude and already proportional,
will be used to derive a proportional from a chosen
R. We also introduce the following parameter for the
resistance against flow:

d
d
d
d
d
d
d
d

kcA
=
H
We assume H to represent the height of each main
basin (in this model both basins are of equal height):
Hz
(1 + z
H
h =
(1 + z
H
h =
(1 + z
Hz
h =
(1 + z

q
c

We introduce the following parameter, which is


loosely defined as one for the pycnal effects:

h =

This turns the flow equation into the following


expression:

S
=

d
c

We further define:

R=

d
d
d
d
d
d
d
d

)
)
)
)

13

1
h
1
=
h
1
=
h
1
=
h
=

1
h
1
=
h
1
=
h
1
=
h

||
(
h
||
(
( ) + ( ) +
h
||
(
( ) + ( ) +
h
||
(
( ) + ( ) +
h
( ) + ( ) +

||
(
h
||
(
( ) + ( ) +
h
||
(
( ) + ( ) +
h
||
(
( x ) + ( ) +
h

( ) + ( ) +

) + ( )
)
)
) + ( )

) + ( )
)
)
) + ( )

In this model, entrainment fluxes are dependent on


the temperature diffusion coefficient as well as the
surface area of the thermocline:
= cAe
w = cAe2
The entrainment in this model may also be
interpreted as a mixing coefficient between the
bottom and top of the two main basins.
Model Initialization
The system of equations is solved numerically
and the time-evolution of the variables in each basin
is plotted. Eventually, a T-S parametric plot for each
basin will show how each basin approaches
equilibrium values for both temperature and salinity
depending on the initial conditions. As in Stommels
initial model, this will allow for examination of the
amount of equilibrium points, their location, whether
certain conditions can make the basins unstable and if
it is salinity- or temperature dominated. As
dimensionless parameters the following are chosen,
partly based on Stommels parameter analysis:
1
6
R=2
1
=
5
e = e = e = 100
1
z =z =z=
5
=

As remaining dimensional parameters we choose:


m
s
kg
= 1027
m
H = 3000m
A = 5 10 m
t = 100
= 20
g = 9.81

Figure 12. Control simulation with symmetrical


initial conditions and high entrainment flow.

Symmetry in the model is directly noticed


because the density lines do not intersect with each
other, as also with Stommels original model. What is
different in the result that is presented is that the
range of initial conditions that lead to the spiral node
is very small (which is only reached from an initial
salinity of 0.5 +/- 0.2). Manipulation of the , R,
parameters leads to loss of equilibria, as Stommel
showed as well. In interpreting the result, freshwater
influx as simulated by a specific initial salinity
condition would lead to hysteresis past the
bifurcation point into the spiral node. Figure 13
shows the time lapse of the THC flow. Comparatively,
if boundary conditions of the real ocean were
perturbed such as to resemble the initial conditions of
the model, the flow is expected to drop to zero.

These last parameters (specifically the last four)


determine the models geometry, so they are left in
their original dimensions. For simplicitys sake, equal
entrainment in the two boxes, as well as equal heights
of the thermocline and surface area are chosen.
Figure 13. Main flow q as a function of time. For
specific initial conditions, flow fluctuates at first,
but returns to zero eventually, simulating a
collapse of the THC.

Results and discussion


A first solution of the model, using symmetry in
the initial conditions to simulate Stommels original
model, shows the same three equilibria, one of which
is unstable (Figure 12).
14

hindering of the natural hysteresis effect as long as


entrainment is high enough. The effect of lower
entrainment flow is shown in Figure 14. Although the
condition of non-intersecting density lines is not met,
the position and occurrence of the equilibria remains
unchanged with a lower entrainment flow.
The second measurable of the model was the
initial condition asymmetry between boxes. For this
last part of the evaluation, a wide range of different
initial conditions were chosen at high entrainment
flow. Each box now has a distinct density profile, but
for the purpose of clarity only that of box one is
shown in Figure 15, the other boxes have the same
bistability. As expected, many of the density lines
overlap and the first time steps show very fast flow,
although the three equilibria are found at exactly the
same positions.
Concluding, it appears that expansion of
Stommels salt-advection box model is not changed in
its core result after complexification by including a
linear entrainment factor and asymmetry in the initial
conditions of the four boxes. This result points
towards further research into modifying the original
model to more accurately represent ocean dynamics.
Parametrisations of processes such as wind stress
could be introduced in order to study the hysteresis
behaviour. In the spectrum of models assessing the
Meridional Overturning Circulation, the model could
be dimensionalised to resemble the conditions in
actual ocean basins.

Figure 14. Entrainment flow of lower orders of


magnitude (Top: e = 10, Bottom: e = 0.1) shows
more chaotic density behaviour in the system, but
the hysteresis result is kept.

5. Conclusion
We set out to examine two main issues of interest
in the study of the Atlantic Meridional Overturning
Circulation, and that of climate in general.
Computational capacity limits a systematic
exploration of the parameter space in AOGCMs over
long timescales, and multi-model ensembles lack a
well-defined metric with which to weight individual
models and interpret the acquired mean. An analysis
of the literature suggests a number of possible
solutions, and the presented model exercise seems to
corroborate these conceptually.
The AMOC is potentially in a bistable regime, and
many modelling efforts to date have mostly
considered surface warming to indicate a linear
decrease in AMOC strength in the 21st century.
Freshwater influx from Greenland ice sheet melt and
river runoff could severely influence this seemingly

Figure 15. Density diagram of the four-box system


with a selection of different initial conditions for
each box. More chaotic than the symmetrical
system, it produces the same equilibria.

In this demonstration case, a choice was made to


have the value of entrainment be at least one order of
magnitude higher than the volume of the basin, so
that its mixing effect is high. In essence, there is no
15

monostable state of the thermohaline mechanism that


drives the circulation in the centuries beyond.
Coupled Atmosphere-Ocean General Circulation
Models represent the pinnacle of detail and accuracy
in past, present and future climate projections. Their
development over the last decades has seen
integration of more physical processes, higher
resolution, and mediation of internal variability and
uncertainty. Nevertheless, their computational cost is
high, which is why the modelling community has
turned to models of lesser complexity in order to
assess feedbacks and long timescale forcing scenarios.
EMICs have been shown to be, at least in coarse
resolution, on par with AOGCMs in their predictive
capacity, and simple models such as Stommels are
clear in capturing specific mechanisms such as
thermohaline hysteresis. Recent freshwater forcing
simulations using EMICs and incidentally even stateof-the-art AOGCMs have sparked a renewed interest
in this hysteresis behaviour.
The great variety in climate models comes with a
need for representative future projections, which is
why multi-model ensembles are used to provide a
realistic consensus of modelling efforts. Various
weighting methods to mediate the relative quality of
ensemble members have been developed, based on
probabilistic intercomparison in the context of
observational data, although an objective metric is yet
to be agreed upon. Perturbed physics ensemble
simulations seem to allow a systematic exploration of
modelling parameter space, although the model
chosen to perturb could have imperfections in the
core, influencing the ensemble in its totality.
The modelling exercise employed in this thesis is
based on continued attention to Stommels saltadvective feedback mechanism. It shows that even
when physical and geometrical expansions to the
original two-box model in the form of thermocline
entrainment and asymmetrical four-box initial
conditions are made, the classical bistability of the
simulated thermohaline circulation is kept. The
results indicate that the original model could be
expanded further to introduce more realistic
geometries and physical processes such as windstress forcing in order to simulate real ocean basins.
We conclude by arguing for a wider scope of
complexity in climate modelling, in order to mitigate
computational cost, study specific climate change
mechanisms in conceptual detail, and examine longterm climate forcing from output of AOGCMs. EMICs
16

in the range close to AOGCMs, and simple climate


models on the conceptual end of the spectrum, can be
vitally important in assessing a variety of forcing
scenarios. The models of lesser complexity might
have less accuracy, but can much more easily be
tuned in their parameters and provide probabilistic
uncertainty distributions. As a final note, the tentative
underappreciation of freshwater influx in some of the
state-of-the-art multi-model ensembles calls for
further research in the thermohaline hysteresis
mechanism, from all sides of the modelling spectrum.
References
Claussen, M., L. A. Mysak, A. J. Weaver, M. Crucifix, T. Fichefet, M. F. Loutre,
S. L. Weber, J. Alcamo, B. A. Alexeev, A. Berger, R. Calov, A. Ganopolski,
H. Goosse, G. Lohmann, F. Lunkeit, I. I. Mokhov, P. Petoukhov, P. Stone,
Z. Wang, 2002. Earth System Models of Intermediate Complexity:
Closing the Gap in the Spectrum of Climate System Models. Climate
Dynamics 18: 579-586.
Collins, M., B. B. B. Booth, B. Bhaskaran, G. R. Harris, J. M. Murphy, D. M. H.
Sexton, M. J. Webb, 2010. Climate Model Errors, Feedbacks and
Forcings: a Comparison of Perturbed Physics and Multi-Model
Ensembles. Climate Dynamics 36: 1737-1766.
Cubasch, U., et al., 1990: Processes and modelling. In: Climate Change: The
IPCC Scientific Assessment [Houghton, J.T., G.J. Jenkins, and J.J.
Ephraums (eds.)]. Cambridge University Press, United Kingdom and
New York, NY, USA, pp. 6991.
Forest, C. E., P. H. Stone, A. P. Sokolov, M. R. Allen, M. D. Webster, 2002.
Quantifying Uncertainties in Climate System Properties with the Use
of Recent Climate Observations. Science: 295, 113117.
Giorgi, F., L. O. Mearns, 2002. Calculation of Average, Uncertainty Range
and Reliability of Regional Climate Changes from AOGCM Simulations
via the Reliability Ensemble Averaging (REA) Method. Journal of
Climate 15: 11411158.
Goldstein, M., and J. C. Rougier, 2004. Probabilistic Formulations for
Transferring Inferences from Mathematical Models to Physical
Systems. Journal of Scientific Computing 26: 467487.
Gregory J. M., K. W. Dixon, R. J. Stouffer, A. J. Weaver. et al., 2005. A Model
Intercomparison of Changes in the Atlantic Thermohaline Circulation
in Response to Increasing Atmospheric CO2 Concentration.
Geophysical Research Letters 32, DOI: 10.1029/2005GL023209.
Guan, T. P., R. X. Huang, 2008. Stommels Box Model of Thermohaline
Circulation RevisitedThe Role of Mechanical Energy Supporting
Mixing and the Wind-Driven Gyration. Journal of Physical
Oceanography 38: 909-918.
Hargreaves, J. C., J. D. Annan, 2004. Using Ensemble Prediction Methods to
Examine Regional Climate Variation under Global Warming Scenarios.
Ocean Modelling 11: 174-192.
Harvey, D., et al., 1997: An Introduction to Simple Climate Models Used in
the IPCC Second Assessment Report. IPCC Technical Paper 2
[Houghton, J.T., L.G. Meira Filho, D.J. Griggs, and K. Maskell (eds.)].
IPCC, Geneva, Switzerland, 51 pp.
Hawkins, E., R. S. Smith, L. C. Allison, J. M. Gregory, T. J. Woollings, H.
Pohlmann, B. De Cuevas, 2011. Bistability of the Atlantic overturning
circulation in a global climate model and links to ocean freshwater
transport.
Geophysical
Research
Letters
38,
DOI:10.1029/2011GL047208.
Hofmann, M., S. Rahmstorf, 2009. On the Stability of the Atlantic
Meridional Overturning Circulation. Proceedings of the National
Academy of Sciences of the United States of America 106: 2058420589.
Intergovernmental Panel on Climate Change, 2007. The Physical Science
Basis. Contribution of Working Group I to the Fourth Assessment

Report of the Intergovernmental Panel on Climate Change. Cambridge


University Press, United Kingdom: Chapters 8 and 10.

Schmittner, A., M. Latif, B. Schneider, 2005. Model Projections of the North


Atlantic Thermohaline Circulation for the 21st Century Assessed by
Observations.
Geophysical
Research
Letters
37,
DOI:10.1029/2005GL024368.

Jackson, L. C., M. Vellinga, G. R. Harris, 2011. The Sensitivity of the


Meridional Overturning Circulation to Modelling Uncertainty in a
Perturbed Physics Ensemble Without Flux Adjustment. Climate
Dynamics, DOI: 10.1007/s00382-011-1110-5.

Schneider, B., M. Latif, A. Schmittner, 2007. Evaluation of Different


Methods to Assess Model Projections of the Future Evolution of the
Atlantic Meridional Overturning Circulation. Journal of Climate 20:
2121-2133.

Kettleborough, J. A., B. B. B. Booth, P. A. Stott, and M. R. Allen, 2007.


Estimates of Uncertainty in Predictions of Global Mean Surface
Temperature. Journal of Climate 20: 843855.

Selten, F. M., G. W. Branstator, M. Kliphuis, H. A. Dijkstra, 2004. Tropical


Origins for Recent and Future Northern Hemisphere Climate Change.
Geophysical Research Letters 31, DOI:10.1029/2004GL020739.

Knutti, R., D. Masson, 2011. Climate Model Genealogy. Geophysical


Research Letters 38, DOI:10.1029/2011GL046864.

Smith, L. A., 2002. What Might We Learn from Climate Forecasts?


Proceedings of the National Academy of Sciences of the United States
of America 99: 24872492.

Knutti, R., R. Furrer, C. Tebaldi, J. Cermak, G. Meehl, 2010. Challenges in


Combining Projections from Multiple Climate Models. Journal of
Climate 23: 2739-2755.

Smith, R. L., C. Tebaldi, D. Nychka, L. O. Mearns, 2009. Bayesian Modeling


of Uncertainty in Ensembles of Climate Models. Journal of the
American Statistical Association 104: 97-117.

Kuhlbrodt, T., A. Griesel, M. Montaya, A. Levermann, M. Hofmann, S.


Rahmstorf, 2007. On the driving processes of the Atlantic meridional
overturning circulation, Reviews of Geophysics 45: 8755-8785.

Stainforth, D. A., T. Aina, C. Christensen, M. Collins, et al., 2005. Uncertainty


in Predictions of the Climate Response to Rising Levels of Greenhouse
Gases. Nature 433: 403406.

Kuhlbrodt, T., K. Zickfeld, S. Rahmstorf, F. B. Vikeb, S. Sundby, M.


Hofmann, P. M. Link, A. Bondeau, W. Cramer, C. Jaeger, 2009. An
Integrated Assessment of Changes in the Thermohaline Circulation.
Climatic Change 96: 489-537.

Stocker, T. F., and A. Schmittner, 1997. Influence of CO2 Emission Rates on


the Stability of the Thermohaline Circulation. Nature: 388, 862865.

Ledwell J. R., Watson A. J., Law C. S., 1993. Evidence for Slow Mixing Across
the Pycnocline from an Open Ocean Tracer Release Experiment.
Nature 364:701703.

Stommel, H., 1961. Thermohaline Convection with Two Stable Regimes of


Flow. Tellus: 31, 224-230.

Manabe, S., and R.J. Stouffer, 1995. Simulation of Abrupt Climate Change
Induced by Fresh Water Input to the North Atlantic Ocean. Nature
378: 165167.

Stouffer, R.J., J. Yin, J. M. Gregory, K. W. Dixon, et al., 2006. Investigating the


Causes of the Response of the Thermohaline Circulation to Past and
Future Climate Changes. Journal of Climate 19: 13651387.

Manabe, S., Bryan, K., 1969. Climate Calculations with a Combined OceanAtmosphere Model. Journal of the Athmospheric Sciences 26: 786789.

Tebaldi, C., J. M. Arblaster, R. Knutti, 2011. Mapping Model Agreement on


Future Climate Projections. Geophysical Research Letters 38,
DOI:10.1029/2011GL049863.

Marotzke, J., 2000. Abrupt Climate Change and Thermohaline Circulation:


Mechanisms and Predictability. Proceedings of the National Academy
of Sciences of the United States of America 97: 1347-1350.

Tebaldi, C., R. L. Smith, D. Nychka, L. O. Mearns, 2004. Quantifying


Uncertainty in Projections of Regional Climate Change: A Bayesian
Approach to the Analysis of Multimodel Ensembles. Journal of Climate
18: 1524-1539.

McAvaney, B.J., et al., 2001: Model evaluation. In: Climate Change 2001:
The Scientific Basis. Contribution of Working Group I to the Third
Assessment Report of the Intergovernmental Panel on Climate Change
[Houghton, J.T., et al. (eds.)]. Cambridge University Press, United
Kingdom and New York, NY, USA, pp. 471523.

Trenberth, K. E., and J. M. Caron, 2001. Estimates of Meridional


Atmosphere and Ocean Heat Transports. Journal of Climate 14: 3433
3443.
Weber SL, S. S. Drijfhout, A. Abe-Ouchi, M. Crucifix, M. Eby, A. Ganopolski,
S. Murakami, B. Otto-Bliesner, W. R. Peltier, 2007. The Modern and
Glacial Overturning Circulation in the Atlantic Ocean in PMIP Coupled
Model Simulations. Climate of the Past 3: 5164.

Meehl, G. A., C. Covey, T. Delworth, M. Latif, B. McAvaney, J. F. B. Mitchell, R.


J. Stouffer, and K. E. Taylor, 2007. The WCRP CMIP3 Multi-Model
Dataset: A New Era in Climate Change Research. Bulletin of the
American Meteorological Society 88: 1383-1394.

Weigel, A. P., R. Knutti, M. A. Liniger, C. Appenzeller, 2010. Risks of


Weighting in Multimodel Climate Projections. Journal of Climate 23:
4175-4189.

Palmer, T. N., G. J. Shutts, R. Hagedorn, F. J. Doblas-Reyes, T. Jung, M.


Leutbecher, 2005. Representing Model Uncertainty in Weather and
Climate Prediction. Annual Review of the Earth and Planetary Sciences
33: 163-193.

Whitehead, J. A., 2008. Abrupt Transitions and Hysteresis in Thermohaline


Laboratory Models. Journal of Physical Oceanography 39: 1231-1244.

Prange, M., G. Lohmann, A. Paul, 2003. Influence of Vertical Mixing on the


Thermohaline Hysteresis: Analyses of an OGCM. Journal of Physical
Oceanography 33: 1707-1721.

Wunsch, C., 2002. What is the thermohaline circulation? Science 298:


1180-1181.

Rahmstorf, S., 2000. The Thermohaline Ocean CirculationA System with


Dangerous Thresholds? Climate Change 46: 247256.
Rahmstorf, S., 2002. Ocean Circulation and Climate During the Past
120,000 Years. Nature 419: 207214.
Rahmstorf S, M. Crucifix, A. Ganopolski, H. Goosse, I. Kamenkovich, R.
Knutti, G. Lohmann, R. Marsh, L. A. Mysak, Z. Wang, A. J. Weaver, 2005.
Thermohaline circulation hysteresis: A Model Intercomparison.
Geophysical Research Letters 32, DOI:10.1029/2005GL023655.
Raper, S.C.B., J.M. Gregory, and T.J. Osborn, 2001. Use of an UpwellingDiffusion Energy Balance Model to Simulate and Diagnose AOGCM
Results. Climate Dynamics 17 : 601613.
Rooth, C., 1982. Hydrology and Ocean Circulation. Progress in
Oceanography, 11, 131149.
Sausen, R., K. Barthel, and K. Hasselmann, 1988. Coupled OceanAtmosphere Models with Flux Correction. Climate Dynamics 2 : 145
163.

17

Vous aimerez peut-être aussi