Introduction
The atmosphere at mid-latitudes displays a variability on a wide range of
space scales and timescales, and in particular a low-frequency variability at
interannual and decadal timescales as suggested by the analyses of different
time series developed in the past years
. In
contrast to the phenomenon of El Niño–Southern Oscillation (ENSO), of
which the driving mechanisms are intensively studied and quite well
understood
Recently the impact of the coupling between the ocean and the atmosphere at
mid-latitudes on the atmospheric predictability
and the development of the LFV has been explored in a series of
low-order coupled ocean–atmosphere systems. However, the limited flexibility
of the possible geometries of these previous models led the present authors
to develop a series of new model versions. The first of these, OA-QG-WS v1
, for Ocean-Atmosphere–Quasi-Geostrophic–Wind Stress,
features only mechanical coupling between the ocean and the atmosphere, and
uses 12 atmospheric variables following and four oceanic modes
following . In a successor of this model, OA-QG-WS v2, the set
of atmospheric variables is extended from 12 to 20 as in . This
increase in resolution in the atmosphere was shown to be key to the
development of a realistic double gyre in the ocean . A third
version of this model, hereafter referred to as
In the
In this article, we present a model that generalizes the
Model formulation
The model is composed of a two-layer quasi-geostrophic (QG) atmosphere, coupled both thermally and mechanically to a QG shallow-water ocean layer, in the -plane approximation. The atmospheric component is an extension of the QG model, first developed by and further refined by . The equations of motion for the atmospheric streamfunction fields at 250 hPa and at 750 hPa, and the vertical velocity , read
The Coriolis parameter is linearized around a value estimated at latitude N, , with . The parameters and quantify the friction between the two atmospheric layers and between the ocean and the atmosphere, respectively, and hPa is the pressure difference between the atmospheric layers.
The equation of motion for the streamfunction of the ocean
layer reads
The time evolution of the atmosphere and ocean temperatures and obeys the following equations: Here, and are the heat capacities of the atmosphere and the active ocean layer. is the atmospheric barotropic streamfunction. is the heat transfer coefficient at the ocean–atmosphere interface, and is the static stability of the atmosphere, taken to be constant. The quartic terms represent the long-wave radiation fluxes between the ocean, the atmosphere, and outer space, with the emissivity of the grey-body atmosphere and the Stefan–Boltzmann constant. By decomposing the temperatures as and , the quartic terms are linearized around spatially uniform temperatures and , as detailed in Appendix B of . and are the short-wave radiation fluxes entering the atmosphere and the ocean that are also decomposed as and .
The hydrostatic relation in pressure coordinates where the geopotential height and the ideal gas relation allow one to write the spatially dependent atmospheric temperature anomaly , with often referred to as the baroclinic streamfunction. is the ideal gas constant. This can be used to eliminate the vertical velocity from Eqs. ()–() and (). This reduces the independent dynamical fields to the streamfunction fields and , and the spatially dependent temperatures and .
The prognostic equations for these four fields are then non-dimensionalized by dividing time by , distance by a characteristic length scale , pressure by the difference , temperature by , and streamfunction by . A more detailed discussion of the model equations and their non-dimensionalization can be found in and .
All the parameters of the model equations used in the present work are listed in Table .
Values of the parameters of the model that are used in the analyses of Sect. .
Parameter (unit) | Value | Parameter (unit) | Value |
---|---|---|---|
(km) | |||
(km) | (kg m) | ||
(s) | (W m K) | ||
(W m K) | (m s Pa) | ||
(s) | (m s) | ||
(s) | (J kg K) | ||
(W m) | (J m K) | ||
(W m) | (J m K) | ||
(s) | (K) | ||
(s) | (K) | ||
(m) |
Expansion of the dynamical fields
In non-dimensionalized coordinates and , the domain is defined by , with the aspect ratio between its meridional and zonal extents (see Table for the value used here). The atmospheric flow is defined in a zonally periodic channel with no-flux boundary conditions in the meridional direction ( at ), whereas the oceanic flow is confined within an ocean basin by imposing no-flux boundaries in both the meridional ( at ) and zonal ( at ) directions. These boundary conditions limit the functions used in the Fourier expansion of the dynamical fields. With the proper normalization, the basis functions for the atmosphere must be of the following form, following the nomenclature of :
Analogously, the oceanic basis functions must be of the form with integer values of , , , , and .
For example, the spectral truncation used by can be specified
as Eqs. ()–() with ; .
extend this set by two blocks of two functions each, and the
resulting set can be specified as ; . The
For the given ranges of , and , and , the number of basis functions can be calculated as
Ordering the basis functions as in Eqs. ()–(), along increasing values of and then , allows one to write the set as . The dynamical fields can then be written as the following truncated series expansions:
Furthermore, the short-wave radiation or insolation is determined by ; . In Eq. (), a term is added to the oceanic basis function in order to give it a vanishing spatial average. This is required to guarantee mass conservation in the ocean , but otherwise does not affect the dynamics. Indeed, it can be added a posteriori when plotting the field . This term is non-zero for odd and , The mass conservation is automatically satisfied for , as the spatial averages of the atmospheric basis functions are zero.
Substituting the fields in Eqs. ()–() and projecting on the different basis functions yield ordinary differential equations (ODEs) for as many variables. Due to the linearization of the quartic temperature fields in Eqs. () and (), these equations are at most bilinear (due to the advection term) in the variables , and , which will henceforth jointly be referred to as , the components of the state vector .
To construct the dynamical equations of these variables, one has to compute the various projections or inner products with the basis functions, for which the following shorthand notation will be used:
As described by , the inner products for the atmosphere can be computed as purely algebraic formulae of the wave numbers , , and . We reiterate these algebraic formulae in Sect. of Appendix and extend them with the formulae for both the ocean–atmosphere coupling terms and the ocean inner products in Sect. . The inner products can be represented as either two-dimensional or three-dimensional tensors, which are sparse but generally not diagonal.
Technical implementation
Substituting the fields by Eqs. ()–() and calculating the coefficients using the expressions for the inner products as in Appendix yields a set of prognostic ordinary differential equations. These equations are at most bilinear in the variables ) due to the linearization of the radiative terms around a reference temperature present in Eqs. ()–(). This system of ODEs can therefore be most generically expressed as the sum of a constant, a matrix multiplication, and a tensor contraction: This expression can be further simplified by adding a dummy variable that is identically equal to one: . This extra variable allows one to merge , , and into the tensor , in which the linear terms are represented by and the constant term by : The elements of the tensor are specified in Appendix . Recasting the system of ordinary differential equations for in the form of a tensor contraction has certain advantages, as we will clarify below. The symmetry of Eq. () allows for a unique representation of , if it is taken to be upper triangular in the last two indices ( if ). Since is known to be sparse, it is stored using the coordinate list representation, i.e. a list of tuples . This representation renders the computation of the tendencies computationally very efficient as well as conveniently parallelizable.
Two implementations of
Derivation of Jacobian, tangent linear, and adjoint models
The form of Eq. () allows one to easily compute the Jacobian matrix of this system of ODEs. Indeed, denoting the right-hand side of Eq. () as , the expression reduces to The differential form of the tangent linear (TL) model for a small perturbation of a trajectory is then simply To obtain the differential form of the adjoint model along the trajectory , the Jacobian is transposed to yield the following equations for the adjoint variable :
Model dynamics
This section details some key results obtained with the model for various levels of spectral truncation, with the set of parameter values given in Table . The parameter values for , , , , , , , , and were selected as detailed in . The same value was chosen for and , as was done in ; see also . Unless otherwise stated, all the following results are obtained after first integrating the model for a transient period of 30 726.5 years. The model is subsequently integrated for another 92 179.6 years to obtain a sufficiently long trajectory from which good statistics can be extracted.
For the atmospheric part of the model, a previous study , referred to as CT in the following, has shown that spurious chaos and a too large variability in the modes near the spectral cut-off could take place if the resolution is not high enough. These manifestations of spurious behaviour can lead to solutions that differ significantly from the solutions of the full partial differential equations (PDEs, here Eqs. –). These findings lead us to the important question of convergence: to what degree has the solution of the truncated equations converged towards the solution of the PDEs? Although we do not have access to the latter, one can infer how the solutions are altered when the resolution is increased. Therefore, it cannot be asserted that convergence has been reached, and this point was also clearly stated in CT. However, we can reasonably suppose that when the solutions stabilize, they give an insight into the full dynamics.
Number of variables, transient time, and effective runtime of the runs (in years).
Resolution | No. of | Transient | Eff. runtime |
---|---|---|---|
variables | () | () | |
atm. – oc. – | 30 726.5 | 92 179.6 | |
atm. – oc. – | 30 726.5 | 92 179.6 | |
atm. – oc. – | 30 726.5 | 92 179.6 | |
atm. – oc. – | 30 726.5 | 92 179.6 | |
atm. – oc. – | 30 726.5 | 92 179.6 | |
atm. – oc. – | 30 726.5 | 92 179.6 | |
atm. – oc. – | 30 726.5 | 92 179.6 | |
atm. – oc. – | 30 726.5 | 92 179.6 | |
atm. – oc. – | 30 726.5 | 92 179.6 | |
atm. – oc. – | 30 726.5 | 92 179.6 | |
atm. – oc. – | 15 363.3 | 92 179.6 | |
atm. – oc. – | 15 363.3 | 74 972.7 |
Cross section of the attractors for various model resolutions. The atmospheric and oceanic resolutions are both indicated above each panel. The parameters are given in Table .
[Figure omitted. See PDF]
Cross section of the attractors for various model resolutions (continued from Fig. ).
[Figure omitted. See PDF]
This question is now addressed for the
Variance distributions of the variables in percents for various model resolutions. For the variables associated with the -type basis functions, the wave numbers and are not defined (nd).
[Figure omitted. See PDF]
The first panel of Fig. , with the atm. – oc. – resolution, shows the typical attractor geometry found in and with a noisy, seemingly periodic orbit associated with the development of a large low-frequency signal. However, as the resolution is increased in both the ocean and atmosphere components, this structure destabilizes and we obtain more compact, noisy attractors in Figs. and . The cause of this structural change is an interesting question in itself, which is worth exploring further in the future, as it is associated with the problem of structural stability of models, but is beyond the scope of the present work.
Regarding the question of convergence, the variability of the atmospheric variables becomes quite stable as the resolution increases beyond –. Indeed, the bounds of the attractors on the vertical axis () stabilize at this resolution. This result is in agreement with the findings of CT. On the other hand, the convergence is not yet reached for the oceanic variables whose variability is strongly affected by adding further modes as in the – and – resolutions.
The impact of the resolution on the solutions can also be examined by computing the variance of each variable of the barotropic and baroclinic streamfunctions, since these are associated with the kinetic and potential energy of the system . The presence of spurious behaviour can then be detected through substantial changes in this variability. The distributions of the total variance of the variables and are depicted in Figs. –. The results show that the variance distribution does not change much beyond the – resolution for the atmospheric component. However, for the oceanic component, the variance distribution is strongly modified when the resolution increases, and therefore one cannot conclude from Fig. that some sort of convergence is reached at the – resolution. To interpret this specific property, one must recall an important feature of two-dimensional quasi-geostrophic turbulence, namely the presence of a specific space scale, the Rhines scale, which delimits the two regimes associated with a wave-dominated dynamics and a turbulent dynamics. This space scale is given by
Variance distributions of the variables in percents for various model resolutions (continued from Fig. ). For the variables associated with the -type basis functions, the wave numbers and are not defined (nd).
[Figure omitted. See PDF]
where represents the root-mean-square velocity of the energy-containing scales and is the meridional derivative of the Coriolis parameter . If one takes the typical velocity of the order of a few metres per second and a few centimetres per second within the atmosphere and the ocean at large scales, the typical length scales will be of the order of 1000 and 100 km, respectively. Therefore the highest wave numbers necessary to resolve the wave-dominated part within the atmosphere and the ocean differ by a factor of 10. Coming back to our analysis, if this limit is reached for the atmosphere in our model at , –5, we should suspect that a value of –50 should be used for the ocean. This of course imposes strong constraints on our reduced-order model and would considerably limit its flexibility.
Variance distributions of the variables in percents for various model resolutions.
[Figure omitted. See PDF]
Let us now focus on the development of the LFV in these different model configurations, and let us define the geopotential height difference between the locations () and () of the model's non-dimensional domain: where is the geopotential height field, as in . The results shown in Figs. and indicate that the LFV, present for atm. – oc. – as in , is a very weak signal at intermediate resolutions, but develops again when the number of modes is increased, as shown by the 1-year and 5-year running means. It suggests that the LFV previously found in low-resolution versions (see Fig. , panel atm. – oc. –) is a robust feature of the model. Moreover, at high resolutions this LFV is weaker than for the VDDG model version, but it seems closer to the actual dynamics found for the North Atlantic Oscillation (NAO) as discussed in and .
Variance distributions of the variables in percents for various model resolutions (continued from Fig. ).
[Figure omitted. See PDF]
Time series of the geopotential height difference (m) between locations ( and ( of the model's non-dimensional domain for different resolutions. Running averages for (black) and (red) are also provided, highlighting the LFV signal present in the series.
[Figure omitted. See PDF]
Time series of the geopotential height difference (continued from Fig. ).
[Figure omitted. See PDF]
The climatologies of the atmospheric barotropic streamfunction expressed in geopotential height further highlight the changes in the statistical properties of the model as a function of resolution. As shown in Figs. and , the convergence is pretty fast toward an averaged zonal atmospheric circulation as the model resolution is increased. By contrast, the convergence for the oceanic streamfunction is less clear (Figs. and ), although a recurrent “global” double gyre is present for each resolution. As for the LFV, the topology of the gyres at high resolutions and their small-scale structures also seem to depend on whether , , and , are even or odd numbers.
The previous results point toward the important question of the optimal resolution of the oceanic component needed to get a sufficiently low-resolution model while keeping a dynamics with strong similarities to a very high-resolution model. To answer this question, we have performed some higher-resolution integrations, but on shorter time spans. The time span for each integration is given in Table .
The variance distributions of the oceanic streamfunction variables (see Fig. ) have decreased at the spectral cut-off's edges compared to the distributions of the lower-resolution model configurations shown in Fig. . However, this decrease is not sufficient, and apparently spurious effects are still present. For instance, the decay is not identical in both directions, with a slower decay rate as the zonal wave number increases. We can even notice a peak in the distribution around and for all these higher model resolutions. This indicates that in fact we are still far from a quantitatively representative solution in the ocean. It confirms that, as stated previously, a resolution of the order of the Rhines scale is needed to achieve a good convergence. For the ocean, it corresponds to a 100 km resolution which would then require roughly 2000 modes. Such a model will of course be very computationally expensive and cannot be considered a “reduced”-order model anymore.
Climatologies for the geopotential height field (m) presented on the non-dimensional model domain, as obtained using 92 179.6 years of model integration.
[Figure omitted. See PDF]
Climatologies for the geopotential height field (m) (continued from Fig. ).
[Figure omitted. See PDF]
However, the comparison between the atm. – oc. – model
configuration and the – or – model configurations shows
that the former displays a large-scale behaviour close to the latter two, but
with a reduced complexity and computational cost. This similarity can be
assessed by considering the climatologies of these higher-resolution runs
displayed in Fig. and by watching the
corresponding videos (see below). We therefore believe that the atm.
– oc. – model configuration is a good candidate when
investigating more realistic dynamics than the one presented in
Climatologies for the oceanic streamfunction field (m s) presented on the non-dimensional model domain, as obtained using 92 179.6 years of model integration.
[Figure omitted. See PDF]
Climatologies for the oceanic streamfunction field (m s) (continued from Fig. ).
[Figure omitted. See PDF]
Finally, the dynamics of the model for the various resolutions are also illustrated in the videos provided as supplementary material. These videos depict the time evolution of the streamfunction and temperature fields, as well as the geopotential height difference and the three-dimensional phase-space projection shown in Figs. and . They give an insight into the high-frequency atmospheric and low-frequency oceanic variability, and also show the interesting time evolution of the oceanic gyres. In these videos, a striking feature is the presence of a westward wave propagation within the ocean while the LFV is developing in the coupled system. This feature has been associated with the propagation of Rossby-like waves .
Conclusions
A new reduced-order coupled ocean–atmosphere model is presented, extending
the low-resolution versions previously published . It
is referred to as
Variance distributions of the variables in percents for the high-resolution runs. The information on the different runtimes is gathered in Table .
[Figure omitted. See PDF]
In the present work, we have studied the impact of the resolution on the model solution's dynamics, by investigating the properties of the attractors and the variance distributions in both the oceanic and atmospheric components. The conclusion that can be drawn is that the convergence of the atmospheric component of the system is quite fast (as noted in ), with variance distributions decreasing rapidly as a function of scale. However, the convergence of the oceanic component is much slower. Consequently, none of the solutions presented so far have satisfactorily converged toward a dynamics that correctly reflects the wave-dominated regime of the coupled ocean–atmosphere system. This regime corresponds to a resolution associated with the Rhines scale (which for the ocean is equal to 100 km or, equivalently, to wave numbers of the order of ). This stresses the need for high-resolution oceanic models to correctly represent the full coupled dynamics. One coupled model configuration which could, however, be recommended so far is the atm. – oc. – configuration, which seems to display some robustness in the ocean climatology as compared to the full – and – configurations. This conclusion requires further investigation with even higher resolutions, together with the use of more advanced tools of analysis like the computation of the Lyapunov exponents as in . These can be computed using the tangent linear model version for which an implementation is also provided. This will be the subject of a future investigation.
Climatologies for the oceanic streamfunction field (m s) presented on the non-dimensional model domain for the high-resolution runs displayed in Fig. .
[Figure omitted. See PDF]
The robustness of the LFV pattern, one of the most interesting features of the model, has also been explored. As it turns out, a LFV is still present in a large portion of the model configurations explored (not in –, –, and –), but a weaker LFV signal is found when high-resolution configurations are used. A dominant signal is found with a wide variety of periods ranging from 1 to 100 years, depending on the model configuration. A more detailed analysis of the underlying structure of the system's attractor is needed to clarify the origin of this diversity, for instance through a bifurcation analysis as in . Note that the VDDG model is still an important tool in this hierarchy of models, since it already contains the basic mechanisms leading to the LFV.
Another interesting finding is the change of structure of the climatologies of the ocean gyres when choosing even or odd wave numbers (, , and , ). Is this feature purely associated with the convergence toward a spatially continuous field, or does it reflect specific properties of the dynamical equations, such as symmetries or invariance? These questions are still open and will be the subject of a future investigation that should allow one to clarify the best set of modes needed for the ocean description.
Finally, the aim of the model is to study the effects of specific physical interaction mechanisms between the ocean and the atmosphere on the mid-latitude climate, both at large and intermediate scales. The modular design of the code of the model is adapted to such purposes, with the possibility of implementing new components, such as oceanic active transport, time-dependent forcings, or salinity fields.
Code availability
Formulae to compute the inner products
In the formulae of the inner products of the atmospheric modes, use the following helper functions: The same notation will be used in this appendix. In what follows, is the Kronecker delta, so that if , and 0 otherwise. Likewise, the function used in this appendix is defined as Using these functions, the various coefficients of the model are calculated, starting with the internal atmosphere coefficients.
Atmospheric coefficients
In the following, we consider the ordering of the basis function used in Eqs. ()–(). For the sake of clarity, we add an extra informative upper index specifying the type of the atmospheric function in the definitions below. However, the inner products are completely defined by the lower indices alone. The atmospheric functions are thus noted: and the oceanic functions
The coefficients
These coefficients correspond to the eigenvalues of the Laplacian operator acting on the spectral expansion basis functions: which are given for each case by
The coefficients
These coefficients are needed to evaluate the contribution of the -terms, and only involve the - and -type base functions. We have that
The coefficients
These coefficients are given by and the non-zero ones are given by where we have used the functions defined at the beginning of this appendix. All the other permutations can be obtained thanks to
The coefficients
These coefficients are given by Therefore we obtain
The coefficients
These coefficients encode the inner products between the atmospheric and oceanic basis functions: which gives
The coefficients
These coefficients are related to the forcing of the ocean on the atmosphere. They are given by the formula where the are given by the eigenvalues of the Laplacian operator acting on the oceanic basis functions (see next section).
Oceanic coefficients
The coefficients
These coefficients are related to the forcing of the atmosphere on the ocean. They are given by
The coefficients
These coefficients identify with the eigenvalues of the Laplacian acting on the oceanic basis functions:
The coefficients
These coefficients are needed to evaluate the contribution of the -terms and are given by
The coefficients
These coefficients are given by with
The coefficients
These coefficients are given by
The coefficients
These coefficients are related to the short-wave radiative forcing of the ocean and are given by
Definition of the tensor
The system of non-dimensionalized ODEs for the model variables is encoded in the model tensor , of which the complete definition is given in this appendix. Tensor elements that are not listed below are equal to zero. To alleviate the notations, we use a shorthand notation for the indices of the different variables, Furthermore, we suppress the upper indices which indicate the atmospheric function types but are otherwise not needed to unambiguously specify the inner products.
Atmosphere equations
The components of the tensor for the atmosphere streamfunction are given by
The atmospheric temperature equations are determined by the tensor elements where we used the non-dimenionalized quantities
Ocean equations
The components of the tensor for the ocean streamfunction are with , , and .
Finally, the equations for the ocean temperature are determined by where the following non-dimensionalized quantities are used:
Information about the Supplement
The time evolution of the model dynamics is illustrated by a set of videos, which are available online at 10.5446/18552, 10.5446/18553, 10.5446/18554, 10.5446/18555, 10.5446/18556, 10.5446/18557, 10.5446/18558, 10.5446/18559, 10.5446/18560, 10.5446/18561, 10.5446/18562, and 10.5446/18563.
The Supplement related to this article is available online at
Acknowledgements
This work is partly supported by the Belgian Federal Science Policy Office under contract BR/121/A2/STOCHCLIM. The figures and videos have been prepared with the Matplotlib software . Edited by: O. Marti Reviewed by: T. Sengul and one anonymous referee
You have requested "on-the-fly" machine translation of selected content from our databases. This functionality is provided solely for your convenience and is in no way intended to replace human translation. Show full disclaimer
Neither ProQuest nor its licensors make any representations or warranties with respect to the translations. The translations are automatically generated "AS IS" and "AS AVAILABLE" and are not retained in our systems. PROQUEST AND ITS LICENSORS SPECIFICALLY DISCLAIM ANY AND ALL EXPRESS OR IMPLIED WARRANTIES, INCLUDING WITHOUT LIMITATION, ANY WARRANTIES FOR AVAILABILITY, ACCURACY, TIMELINESS, COMPLETENESS, NON-INFRINGMENT, MERCHANTABILITY OR FITNESS FOR A PARTICULAR PURPOSE. Your use of the translations is subject to all use restrictions contained in your Electronic Products License Agreement and by using the translation functionality you agree to forgo any and all claims against ProQuest or its licensors for your use of the translation functionality and any output derived there from. Hide full disclaimer
© 2016. This work is published under http://creativecommons.org/licenses/by/3.0/ (the “License”). Notwithstanding the ProQuest Terms and Conditions, you may use this content in accordance with the terms of the License.
Abstract
This paper describes a reduced-order quasi-geostrophic coupled ocean–atmosphere model that allows for an arbitrary number of atmospheric and oceanic modes to be retained in the spectral decomposition. The modularity of this new model allows one to easily modify the model physics. Using this new model, coined the “Modular Arbitrary-Order Ocean-Atmosphere Model” (
You have requested "on-the-fly" machine translation of selected content from our databases. This functionality is provided solely for your convenience and is in no way intended to replace human translation. Show full disclaimer
Neither ProQuest nor its licensors make any representations or warranties with respect to the translations. The translations are automatically generated "AS IS" and "AS AVAILABLE" and are not retained in our systems. PROQUEST AND ITS LICENSORS SPECIFICALLY DISCLAIM ANY AND ALL EXPRESS OR IMPLIED WARRANTIES, INCLUDING WITHOUT LIMITATION, ANY WARRANTIES FOR AVAILABILITY, ACCURACY, TIMELINESS, COMPLETENESS, NON-INFRINGMENT, MERCHANTABILITY OR FITNESS FOR A PARTICULAR PURPOSE. Your use of the translations is subject to all use restrictions contained in your Electronic Products License Agreement and by using the translation functionality you agree to forgo any and all claims against ProQuest or its licensors for your use of the translation functionality and any output derived there from. Hide full disclaimer