Skip to main content

Advertisement

Log in

Revisiting meridional overturing bistability using a minimal set of state variables: stochastic theory

  • Published:
Climate Dynamics Aims and scope Submit manuscript

Abstract

A stochastic analytical model of the Atlantic meridional overturning circulation (AMOC) is presented and tested against climate model data. AMOC stability is characterised by an underlying deterministic differential equation describing the evolution of the central state variable of the system, the average Atlantic salinity. Stability of an equilibrium implies that infinitesimal salinity perturbations are damped, and violation of this requirement yields a range of unoccupied salinity states. The range of states is accurately predicted by the analytical model for a coupled climate model of intermediate complexity. The introduction of climatic noise yields an equation describing the evolution of the probability density function of the state variable, and therefore the AMOC. Given the hysteresis behaviour of the steady AMOC versus surface freshwater forcing, the statistical model is able to describe the variability of the AMOC based on knowledge of the variability in the forcing. The method accurately describes the wandering between AMOC-On and AMOC-Off states in the climate model. The framework presented is a first step in relating the stability of the AMOC to more observable aspects of its behaviour, such as its transient response to variable forcing.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9

Similar content being viewed by others

References

  • Broecker WS (1991) The great ocean conveyor belt. Oceanography 4:79–87

    Article  Google Scholar 

  • Cunningham SA, Kanzow T, Rayner D et al (2007) Temporal variability of the Atlantic meridional overturning circulation at 26.5° n. Science 317:935–938

    Article  Google Scholar 

  • Dansgaard W, Johnsen SJ, Clausen HB (1993) Evidence for general instability of past climate from a 250-kyr ice-core record. Nature 364:218–220

    Article  Google Scholar 

  • Fuhrer K (1999) Timescales for dust variability in the greenland ice core project (grip) ice core in the last 100,000 years. J Geophys Res 104. doi:10.1029/1999JD900929

  • Gardiner CW (1996) Handbook of stochastic methods for Physics, Chemistry, and the natural sciences. Springer, New York, 442 pp

  • Gregory JM, Dixon K, Stouffer R et al (2007) A model intercomparison of changes in the Atlantic thermohaline circulation in response to increasing atmospheric CO2 concentration. Geophys Res Lett 32. doi:10.1029/2005GL023209

  • Gregory JM, Saenko OA, Weaver AJ (2003) The role of the Atlantic freshwater balance in the hyteresis of the meridional overturning circulation. Clim Dyn 21:707–717. doi:10.1007/s00382-003-0359-8

    Article  Google Scholar 

  • Heinrich H (1988) Origin and consequences of cyclic ice rafting in the northeast Atlantic Ocean during the past 130,000 years. Quat Res 29:142–152

    Article  Google Scholar 

  • Hemming SR (2004) Heinrich events: massive late pleistocene Detritus layers of the North Atlantic and their global climate imprint. Rev Geophys 42:RG1005. doi:10.1029/2003RG000128

    Article  Google Scholar 

  • Hu A, Meehl GA, Han W et al (2012) Role of the Bering Strait on the hysteresis of the ocean conveyor belt circulation and glacial climate stability. Proc Nat Acad Sci 109:6417–6422

    Article  Google Scholar 

  • Livina VN, Kwasniok F, Lenton TM (2010) Potential analysis reveals changing number of climate states during the last 60 kyr. Clim Past 6:77–82

    Article  Google Scholar 

  • Lucarini V, Faranda D, Willeit M (2012) Bistable systems with stochastic noise: virtues and limits of effective one-dimensional Langevin equations. Nonlinear Process Geophys 19:9–22

    Article  Google Scholar 

  • Manabe S, Stouffer RJ (1988) Two stable equilibria of a coupled ocean-atmosphere model. J Clim 1:841–866

    Article  Google Scholar 

  • Manabe S, Stouffer RJ (1999) Are two modes of thermohaline circulation stable? Tellus 51:400–411

    Article  Google Scholar 

  • McManus JF, Francois R, Gherardi JM, Keigwin LD, Brown-Leger S (2004) Collapse and rapid resumption of atlantic meridional circulation linked to deglacial climate changes. Nature 428:834–837

    Article  Google Scholar 

  • Monahan AH (2002) Correlation effects in a simple stochastic model of the thermohaline circulation. Stoch Dyn 2:437–462

    Article  Google Scholar 

  • Monahan AH, Alexander J, Weaver AJ (2008) Stochastic models of the meridional overturning circulation: time scales and patterns of variability. Philos Trans R Soc 366:2525–2542

    Article  Google Scholar 

  • Monahan AH, Culina J (2011) Stochastic averaging of idealised climate models. J Clim 24:3068–3088

    Article  Google Scholar 

  • North Greenland Ice Core Project Members: (2004) High-resolution record of northern hemisphere climate extending into the last interglacial period. Nature 431:147–151

    Article  Google Scholar 

  • Nunes F, Norris RD (2006) Abrupt reversal in ocean overturning during the Paleocene/Eocene period. Nature 439:60–63

    Article  Google Scholar 

  • Pacanowski R (1995) MOM2 documentation user’s guide and reference manual: GFDL Ocean group technical report 3. NOAA, GFDL Princeton, 3rd edn, 232 pp

  • Pardaens AK, Banks HT, Gregory JM (2003) Freshwater transports in hadcm3. Clim Dyn 21:177–195

    Article  Google Scholar 

  • Rahmstorf S (1995) Bifurcations of the Atlantic thermohaline circulation in response to changes in the hydrological cycle. Nature 378:145–149

    Article  Google Scholar 

  • Rahmstorf S (1996) On the freshwater forcing and transport of the Atlantic thermohaline circulation. Clim Dyn 12:799–811

    Article  Google Scholar 

  • Rahmstorf S (2002) Ocean circulation and climate during the past 120,000 years. Nature 419:207–214

    Article  Google Scholar 

  • Rooth C (1982) Hydrology and ocean circulation. Prog Ocean 11:131–149

    Article  Google Scholar 

  • Saenko OA, Weaver AJ (2003) Atlantic deep circulation controlled by freshening in the Southern Ocean. Geophys Res Lett 30. doi:10.1029/2003GL017681

  • Schmittner A, Yoshimori M, Weaver AJ (2002) Instability of glacial climate in a model of the ocean-atmosphere-cryosphere system. Science 295:1489–1493

    Article  Google Scholar 

  • Sijp WP (2012) Characterising meridional overturning bistability using a minimal set of state variables. Clim Dyn. doi:10.1007/s00382-011-1249-0

  • Sijp WP, England MH (2004) Effect of the Drake Passage throughflow on global climate. J Phys Oceanogr 34:1254–1266

    Article  Google Scholar 

  • Sijp WP, England MH, Huber M (2011) Effect of the deepening of the Tasman Gateway on the global ocean. Paleoceanography 26. doi:10.1029/2011PA002143

  • Stommel H (1961) Thermohaline convection with two stable regimes of flow. Tellus 13:224–230

    Article  Google Scholar 

  • Sura P, Newman M, Penland C, Sardeshmukh P (2005) Multiplicative noise and non-Gaussianity: a paradigm for atmospheric regimes?. J Atmos Sci 62:1391–1409

    Article  Google Scholar 

  • Sura P, Penland C (2002) Sensitivity of a double-gyre ocean model to details of stochastic forcing. Ocean Model 4:327–354

    Article  Google Scholar 

  • Vellinga M, Wood RA (2002) Global climatic impacts of a collapse of the Atlantic thermohaline circulation. Clim Change 54:251–267

    Article  Google Scholar 

  • Via R, Thomas D (2006) Evolution of Atlantic thermohaline: early oligocene circulation onset of deep-water production in the North Atlantic. Geology 34:441–444

    Article  Google Scholar 

  • Weaver AJ, Eby M, Wiebe EC et al (2001) The UVic Earth system climate model: model description, climatology, and applications to past, present and future climates. Atmos Ocean 39:1067–1109

    Google Scholar 

  • Weaver AJ, Hughes TMC (1994) Rapid interglacial climate fluctuations driven by North Atlantic ocean circulation. Nature 367:447–450

    Article  Google Scholar 

  • Wei W, Lohmann G, Dima M (2012) Distinct modes of internal variability in the global meridional overturning circulation associated with Southern Hemisphere westerly winds. J Phys Oceanogr 42:785–801

    Article  Google Scholar 

  • Yang S, Galbraith E, Palter J (2013) Coupled climate impacts of the Drake Passage and the Panama Seaway. Clim Dyn. doi:10.1007/s00382-013-1809-6

  • Zahn R, Schoenfeld J, Kudrass HR et al (1997) Thermohaline instability in the North Atlantic during meltwater events: Stable isotope and ice-rafted detritus records from core SO75-26KL, Portuguese margin. Paleoceanography 12. doi:10.1029/97PA00581

Download references

Acknowledgments

We thank the University of Victoria staff for support in usage of their coupled climate model. This research was supported by the Australian Research Council including the ARC Centre of Excellence in Climate System Science. We thank Adam Monahan and an anonymous reviewer for very constructive comments. Willem Sijp was supported by Australian Research Council grant DP1096144.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Willem P. Sijp.

Appendices

Appendix 1: Calculation of diffusivity κ

In Sect. 4 we add a stochastic salt flux H rnd to the surface salt flux of the numerical model and developed a stochastic version of the deterministic theory, involving a Fokker Planck equation, Eq. 4, described in Sect. 3 to determine the validity of that theory. The equation requires an estimate of the effective diffusivity associated with the spreading effect of H rnd on the state S. Here, we present an attempt at calculating this value based on the magnitude of H rnd by considering the smearing effect of the stochastic salt flux H rnd on S. We note that Monahan and Culina (2011) offer a systematic method of stochastic averaging, not used here, to mathematically express the effect of the fast variables on the evolution of the longer time scale variables and that works also with non-Gaussian and auto-correlated noise.

The random values employed in H rnd have a uniform distribution between limits −L < H rnd  < L, where we have chosen L = 34.0 × 10−8 kg/m2/s in the numerical model (all model salt fluxes are in units of kg of salt per m2 per s). This choice is to allow AMOC state transitions but retain distinguishable peaks for the two states. H rnd is updated every 10 years by drawing a new random number, and is added uniformly in time and space over each entire period of 10 years and over the entire Atlantic surface. This is to cause minimal disruption to the ocean model, allow smooth numerical integration and to keep calculation times low in our long model integrations.

To consider the problem of associating a diffusivity κ in S-space with H rnd , we will consider the immediate net effect of the application of H rnd on the change in S in isolation from the long time scale responsive forcing encapsulated in F(S) that comes into play after S has changed more significantly after successive 10 year intervals where H rnd is updated. We use the resulting κ value as the diffusivity in the Fokker-Planck equation, Eq. 4 in the main text.

Recall that we consider the Atlantic basin to include the Arctic and bounded only at approximately 30° south, as Bering Strait is closed in our model. The cumulative change \(\Updelta S_i\) in S due to H rnd alone from the beginning to the end of each 10 year interval, labelled with index i, is \(\Updelta S_i = \frac{A}{V}(H_{rnd})\times10\times365\times24\times3,600\,\hbox{kg}/\hbox{m}^3.\) Here, A = 7.765 × 1013 m2 is the Atlantic surface area and V = 2.492 × 1017 m3 is the Atlantic volume in the numerical model. However, during the 10 year period only a portion 0 < α < 1 of the anomaly \(\Updelta S_i\) remains in the Atlantic as some of it has leaked out of the basin. The net combined effect is an effective change in S of \(\alpha\Updelta S_i.\)

To associate a diffusivity κ with the stochastic perturbations in the numerical model, we examine the discrete time development of S at 10 year time steps t i for \({i \in {\mathbb{Z}}.}\) These new time steps greatly exceed the numerical model time step but are small compared to the millennial time scale of the AMOC processes under investigation (see Sect. 4) They coincide with each time the stochastic perturbation value is updated with a new random number and the state S has attained the value S i . With this interpretation and choice of step sizes, the isolated system exhibits a random walk of S i taking steps of varying sizes \(\alpha\Updelta S_i.\) Approximating the fraction α by a constant value (see below), the step sizes \(\alpha\Updelta S_i\) are uniformly distributed between \(\pm \frac{\alpha AL}{V} \times10\times365\times24\times3,600\,\hbox{kg}/\hbox{m}^3.\) The standard deviation σ of the distribution of each \(\Updelta S_i\) is \(\sigma = 3^{-\frac{1}{2}} \frac{\alpha AL}{V} \times10\times365\times24\times3,600\,\hbox{kg}/\hbox{m}^{3}/10\,\hbox{year}.\)

We can calculate an effective diffusivity κ to help approximate the average squared displacement during the random walk at 10 year step intervals and with varying step sizes in S i . It is related to the expected displacement of S after i steps (assuming no displacement at i = 0 where we choose S 0 = 0) via <S 2 i  >  = 2κ t i . Here, κ has units (kg/m3)2/10 year as it represents diffusion in S-space and we model the system at 10 year time steps. However, as the values of H rnd are uncorrelated at the 10 year steps, <S 2 i > also equals σ2 t i as it is the sum of the variances of the independent changes in S at each step. Therefore, κ = σ2/2 with units (kg/m3)2/10 year. In units of (kg/m3)2/s, diffusion κ in S space arising from H rnd is \(\frac{1}{2} 10\times365\times24\times3,600\,\sigma^2 = 5.9\times10^{-13} \alpha^2.\)

In order to roughly estimate the fraction α of a typical anomaly that remains in the Atlantic for all simulations, we have performed a 10 year experiment where a passive tracer is added uniformly at the Atlantic surface in an identical manner and magnitude to a typical addition of H in the DRY configuration of our numerical model only (and without stochastic forcing). Here, we examine the fraction of the time rate of change in average Atlantic passive tracer concentration due to the passive tracer surface flux over its total rate of change (Fig. 10). The cumulative fraction of passive tracer that remains in the Atlantic in our numerical model simulation is α = 0.89, yielding an effective S change of \(0.89\Updelta S_i.\) Therefore, the effective diffusivity κ for the non-isolated case must be κ = 0.892 × 5.9 × 10−13 (kg/m3)2/s = 0.47 × 10−13 (kg/m3)2/s. This is the value we have used for κ to generate the figures in the main text.

Fig. 10
figure 10

Time series of the rate of (time) change of the average Atlantic passive tracer concentration divided by the rate of change expected from the passive tracer surface flux alone. This is used to estimate the fraction of tracer remaining inside the basin after a 10 year period

In the absence of the longer timescale dynamics encapsulated by F H (S), the application of H rnd would yield a diffusive process acting on ψ(S) with the diffusion coefficient κ calculated above, where S becomes increasingly delocalised and no steady ψ(S) is attained. Incorporating longer time scale oceanic adjustments that act on larger salinity changes, attempting to maintain S near its equilibrium values in the face of the above described random walk tendency (e.g. an AMOC adjustment responding to a net excursion in S arising from the cumulative effect of H rnd ), yields the evolution equation for ψ(S) (Eq. 4) and elucidates how the state S remains in the vicinity of the attractors (AMOC-On and AMOC-Off).

Appendix 2: The glacial change in the disequilibrium flux

The glacial experiments require the application of a lower constant additional fresh water flux H to the climate model to maintain the same salinity states S compared to their Holocene counterparts (Fig. 2). This is equivalent to a lower disequilibrium salt flux \(F_{H_0}(S)\) in the glacial experiments. To briefly examine this further, we will compare the Atlantic budget in the glacial to the Holocene. We shall decompose the Atlantic salt divergence minus the anomalous salt flux associated with H. First, the overturning salt transport component flux F m (e.g. see Rahmstorf 1996) associated with the AMOC, sometimes referred to as the zonal component, is the vertical integral of the zonally integrated velocity times the zonally averaged salinity, evaluated at 32 °S. In addition, the subtropical gyre in the South Atlantic is responsible for a certain salt transport F g across 32 °S, the azonal component. This is simply calculated by subtracting F m from the total ocean salt import, F ocn (a term that is easily diagnosed from models). The total oceanic salt divergence (salt flux into the Atlantic) is then F ocn  = F g  + F m . Finally, the Atlantic salt budget contains a term associated with fresh water exchange at the air-sea interface, −F atm . These components are thoroughly described in Sijp (2012), and are shown for our experiments in Fig. 11.

Fig. 11
figure 11

Decomposition of the Atlantic salt divergence minus the anomalous salt flux associated with H into components related to the gyre, F g (dots), the AMOC, F m (crosses) and the air-sea interface, −F atm (triangles). Note that F atm appears with a minus sign, as F atm denotes a fresh water flux into the ocean. Glacial climate model equilibria are indicated in blue, and the Holocene in red

The glacial F m is reduced with respect to the Holocene in the AMOC-On states (branch on the right), whereas the descrepancy in this term is smaller and reversed in sign for the AMOC-Off states (left branch). In contrast, the glacial NADW outflow as a function of S is greater than in the Holocene in the AMOC-On states (figure not shown). This should not be confused with the outflow for a given H (i.e. the glacial and Holocene outflow in experiments with the same value of H), where the glacial rate is less. In contrast to what we find, the increased glacial outflow for a given S would suggest an increased F m . But this holds only if the weighted vertical salinity gradient at 32°S remains unchanged (Sijp 2012). Therefore, it is the interaction between the spatial distribution of Atlantic salinity and the AMOC that differs between the glacial and Holocene experiments so as to reduce F m in the glacial.

In addition to the glacial AMOC having a more freshening influence on the Atlantic for a given S, the glacial Atlantic is less evaporative than the Holocene, particularly for the AMOC-Off states, as −F atm is positive in both experiments and greater in the Holocene. This may be due to the cooler Atlantic conditions that prevail in the glacial experiments (figures not shown). The gyre term remains relatively similar for the glacial and Holocene climates in the AMOC-On states. In the AMOC-Off states this term is somewhat greater in the glacial. As a result, F glac atm (S) < F Hol ocn (S) and F glac atm (S) > F Hol atm (S) (where the superscript “glac” indicates the glacial and “Hol” the holocene flux). Using F ocn (S) − F atm (S) = H (Eq. 1), we find that a lower fresh water flux H for S in equilibrium with H is required to to maintain the same glacial average salinity S as in the Holocene experiments. As a result, the glacial curve lies below the Holocene curve. This is indeed what we see in Fig. 2. In summary, smaller values of the fresh water flux H are required to maintain identical salinity states S in the glacial. This is because the Atlantic in the glacial experiments is less evaporative, particularly in the AMOC-Off states, and the different interaction between the AMOC and the spatial salinity distribution yields a reduced net influx of salt.

Rights and permissions

Reprints and permissions

About this article

Cite this article

Sijp, W.P., Zika, J.D., d’Orgeville, M. et al. Revisiting meridional overturing bistability using a minimal set of state variables: stochastic theory. Clim Dyn 43, 1661–1676 (2014). https://doi.org/10.1007/s00382-013-1992-5

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00382-013-1992-5

Keywords

Navigation