The Atmospheric Response to a Thermohaline Circulation Collapse: Scaling Relations for the Hadley Circulation and the Response in a Coupled Climate Model

Sybren S. Drijfhout Royal Netherlands Meteorological Institute (KNMI), De Bilt, Netherlands

Search for other papers by Sybren S. Drijfhout in
Current site
Google Scholar
PubMed
Close
Full access

Abstract

The response of the tropical atmosphere to a collapse of the thermohaline circulation (THC) is investigated by comparing two 5-member ensemble runs with a coupled climate model (CCM), the difference being that in one ensemble a hosing experiment was performed. An extension of the Held–Hou–Lindzen model for the Hadley circulation is developed to interpret the results. The forcing associated with a THC collapse is qualitatively similar to, but smaller in amplitude than, the solstitial shift from boreal summer to winter. This forcing results from reduced ocean heat transport creating an anomalous cross-equatorial SST gradient. The small amplitude of the forcing makes it possible to arrive at analytical expressions using standard perturbation theory. The theory predicts the latitudinal shift between the Northern Hemisphere (NH) and Southern Hemisphere (SH) Hadley cells, and the relative strength of the anomalous cross-equatorial Hadley cell compared to the solstitial cell. The poleward extent of the Hadley cells is controlled by other physics. In the NH the Hadley cell contracts, while zonal velocities increase and the subtropical jet shifts equatorward, whereas in the SH cell the opposite occurs. This behavior can be explained by assuming that the poleward extent of the Hadley cell is determined by baroclinic instability: it scales with the inverse of the isentropic slopes. Both theory and CCM results indicate that a THC collapse and changes in tropical circulation do not act in competition, as a possible explanation for abrupt climate change; they act in concert.

Corresponding author address: Dr. Sybren S. Drijfhout, P.O. Box 201, Royal Netherlands Meteorological Institute (KNMI), 3730 AE De Bilt, Netherlands. Email: drijfhou@knmi.nl

Abstract

The response of the tropical atmosphere to a collapse of the thermohaline circulation (THC) is investigated by comparing two 5-member ensemble runs with a coupled climate model (CCM), the difference being that in one ensemble a hosing experiment was performed. An extension of the Held–Hou–Lindzen model for the Hadley circulation is developed to interpret the results. The forcing associated with a THC collapse is qualitatively similar to, but smaller in amplitude than, the solstitial shift from boreal summer to winter. This forcing results from reduced ocean heat transport creating an anomalous cross-equatorial SST gradient. The small amplitude of the forcing makes it possible to arrive at analytical expressions using standard perturbation theory. The theory predicts the latitudinal shift between the Northern Hemisphere (NH) and Southern Hemisphere (SH) Hadley cells, and the relative strength of the anomalous cross-equatorial Hadley cell compared to the solstitial cell. The poleward extent of the Hadley cells is controlled by other physics. In the NH the Hadley cell contracts, while zonal velocities increase and the subtropical jet shifts equatorward, whereas in the SH cell the opposite occurs. This behavior can be explained by assuming that the poleward extent of the Hadley cell is determined by baroclinic instability: it scales with the inverse of the isentropic slopes. Both theory and CCM results indicate that a THC collapse and changes in tropical circulation do not act in competition, as a possible explanation for abrupt climate change; they act in concert.

Corresponding author address: Dr. Sybren S. Drijfhout, P.O. Box 201, Royal Netherlands Meteorological Institute (KNMI), 3730 AE De Bilt, Netherlands. Email: drijfhou@knmi.nl

1. Introduction

It is well established that a collapse of the Atlantic thermohaline circulation (THC) leads to significant cooling over the North Atlantic and northwestern Europe (e.g., Manabe and Stouffer 1994; Vellinga and Wood 2002). The oceanic forcing of the atmosphere occurs through decreased sea surface temperature (SST), especially in regions where convective activity ceases to occur (Vellinga and Wood 2002), and the increase of sea ice coverage. As a result, the atmospheric temperature anomaly peaks in the subpolar North Atlantic and Arctic, suggesting a locally forced heat flux response.

However, the global extent of (pre)historical abrupt climate change necessitates significant changes in tropical climate. And indeed, the atmospheric response to a THC collapse is far reaching (Vellinga and Wood 2002; Zhang and Delworth 2005), that is, well into the tropics, although the signal in SST and surface air temperature is much weaker there. Assuming that Bjerknes compensation occurs for ocean heat transport changes that result from a THC collapse, Seager and Battisti (2007) argued for a tropical source for abrupt climate change that was independent of the THC. Their theory was built on a study of Lee and Kim (2003), in which the relation between tropical SST and latitude and strength of the subtropical jet and midlatitude eddy-driven jet was investigated, suggesting hysteresis behavior and multiple equilibria for these jets. Inspired by this theory, van der Schrier et al. (2007) demonstrated in a simple coupled climate model (CCM) how (sub)tropical SST anomalies indeed force changes in the midlatitude eddy-driven jet.

While smaller fluctuations in meridional overturning circulation (MOC) imply ocean and atmosphere heat transport changes that are subject to the Bjerknes compensation (Shaffrey and Sutton 2006; van der Swaluw et al. 2007), for larger fluctuations the Bjerknes compensation no longer applies and a rearrangement of the earth’s energy budget is necessary to accommodate the change in net meridional energy transport by the coupled ocean–atmosphere system. This holds true in particular for a THC collapse (Vellinga and Wu 2008). Consequently, Vellinga and Wu not only found changes in surface radiation balance and top of the atmosphere (TOA) radiative fluxes that extend into the tropic belt and even into the Southern Hemisphere (SH), they also noticed a significantly changed Hadley circulation. This suggests that the Lee and Kim (2003) theory and the THC theory for abrupt climate change act together. A link between the two could be provided by a theory for the response of the Hadley circulation to a THC collapse.

An important first step toward a theory for the Hadley circulation was taken by Schneider (1977) and Held and Hou (1980, hereafter HH80). They developed a theory that was built on angular momentum conservation. The axisymmetric model predicts, among others, the width of the Hadley cell, the meridional transport, and the distribution of the zonal velocity. Not all aspects of the Hadley circulation were quantitatively described, but the most important features appeared qualitatively correct. Since HH80, the theory for the Hadley circulation was extended further to account for, for example, asymmetric heating and seasonality (Lindzen and Hou 1988, hereafter LH88), narrowness of the heating branch (Hou and Lindzen 1992), moist convection (Fang and Tung 1996), and baroclinic eddy fluxes (Held 2000; Walker and Schneider 2005, 2006).

Because the large-scale tropical temperature response to a THC collapse mainly consists of a dipole pattern associated with reduced cross-hemispheric heat transport (Fig. 1, see also Vellinga and Wood 2002), the theory of LH88 is relevant for this case. Figure 1 shows the heating and cooling pattern associated with a THC collapse comparing two 5-member ensembles of the Ensemble Simulations of Extreme Weather Events under Nonlinear Climate Change (ESSENCE) project (Sterl et al. 2008). The heating pattern is antisymmetric, with respect to the equator, and can be compared, albeit smaller in amplitude, with the differential heating associated with the seasonal cycle. In Fig. 2, a schematic of the asymmetric Hadley circulation is depicted that is consistent with this signal.

One of the main results of LH88, namely, that seasonally varying heating amplifies the annually averaged circulation, has been criticized a number of times. Both a lack of stationarity (Fang and Tung 1999) and vertical diffusion of momentum and eddy fluxes (Walker and Schneider 2005) counteract this amplification, making it doubtful whether the averaged summer and winter solutions differ markedly from the solution for an annually averaged forcing. Because eddies affect the boundary conditions at the poleward edge of the Hadley cell (Walker and Schneider 2006), it can be argued that they compromise the axisymmetric theory for predicting other aspects of the circulation as well. However, changes in relative intensity of the Hadley cells and the latitude of the surface mass flux convergence are predicted by the axisymmetric theory. Also, an eddy-permitting model was shown to recover the sensitivity of the Hadley cell to displacements of the latitude of maximum heating, as predicted by LH88, although it did not reproduce the nonlinear amplification of the annually averaged circulation when seasonal forcing is allowed (Walker and Schneider 2005).

The LH88 framework does not correctly predict the changes in the poleward extent of the Hadley cells and the intensity of the subtropical jets. The changes at the edge of the Hadley cells are controlled by eddy fluxes and can be explained from baroclinic instability considerations (Held 2000) and results from eddy-permitting nonaxisymmetric models (e.g., Walker and Schneider 2005, 2006). This point will be addressed in section 5c. However, other aspects of the circulation can be described within the LH88 framework. Two simplifications to the LH88 model will be introduced that allow for an analytical solution for the response of the Hadley cell to weak asymmetric heating. The first simplification is the small-angle approximation that was already introduced by HH80. The second approximation is the application standard perturbation theory in which the relation between all other variables and the latitude of maximum heating is established. Together, these relations show the intricate link between changes in the THC and changes in the Hadley circulation.

2. The equal-area solution for the Hadley circulation

HH80 and LH88 constructed equal-area solutions for the inviscid Hadley circulation, with symmetric and asymmetric heating, respectively. Their models consist of a few basic elements, which are a thermally driven flow, conservation of angular momentum, geostrophic balance, and energy conservation. The thermal forcing is represented by Newtonian cooling as
i1520-0442-23-3-757-e1
where θ is the vertically averaged potential temperature, θe is the radiative equilibrium temperature, and τ is the radiative relaxation time. The general expression for θe that allows for an off-equatorial heating maximum is given by Eq. (1b) in LH88. For vertically averaged temperature this becomes
i1520-0442-23-3-757-e2
where ϕ0 is the latitude of maximum heating, ΔH is the fractional change in potential temperature from equator to pole, and θ0 is the globally averaged value of θe.
Using the small-angle approximation this equation can be rewritten as
i1520-0442-23-3-757-e3
The antisymmetric heating term can be interpreted as a heating anomaly arising from the temperature difference between the North and South Poles (James 1994),
i1520-0442-23-3-757-e4
Here, ΔNS is defined as the temperature difference between the two poles. The latitude of maximum heating must then obey
i1520-0442-23-3-757-e5
When we define ΔTC as the temperature difference between the Tropics of Capricorn and Cancer, ϕ0 reappears as a scaled temperature difference,
i1520-0442-23-3-757-e6
Now we consider the case that off-equatorial heating can be due to stationary SST anomalies. The diabatic forcing becomes
i1520-0442-23-3-757-e7
We assume that the same radiative time scale applies to SST anomalies as to radiative temperature anomalies. This is true when the SST anomalies are large scale and stationary. Whether this assumption is correct will be tested in section 5, when the appropriate scaling relations are applied to the model data. Here, θf can then be written as
i1520-0442-23-3-757-e8
For certain conditions, the equation for θf becomes similar to the equation for θe, namely, when the SST anomaly consists of a cross-equatorial SST gradient that linearly scales with ϕ. In that case,
i1520-0442-23-3-757-e9
and for ϕ*0 Eq. (6) similarly holds, which defines the off-equatorial maximum heating latitude. In that case, the forcing temperature θf can be written as θe in Eq. (3). However, now ϕ0 is the shift in maximum heating latitude resulting from either changes in ocean circulation, or the seasonality in insolation. The implication of this similarity is that changes in cross-equatorial heat and mass transport in the ocean impose a similar (but smaller) forcing as the insolation change between summer and winter.
To arrive at a model for the response of the Hadley circulation to such SST anomalies, we proceed as in LH88. In the small-angle limit the actual temperature profile is given by
i1520-0442-23-3-757-e10
[see expression (7) in LH88], where ϕ1 is the dividing line between the summer and winter cell (see Fig. 2). The expression for θf can be reformulated to depend on ϕ1 as
i1520-0442-23-3-757-e11
By substituting ϕ0 = ϕ1 = 0 in Eqs. (10) and (11), the expressions (14a) and (14b) from HH80 are recovered.
The equal-area solutions in HH80 and LH88 result from demanding
i1520-0442-23-3-757-e12
and
i1520-0442-23-3-757-e13
Similar equations hold for the SH Hadley cell, with ϕW being replaced by ϕS.
For ϕ0 = ϕ1 = 0, Eqs. (12) and (13) can be solved exactly (in this case ϕW = −ϕS = ϕH), giving a scaling relation for ϕH. When ϕ0 and ϕ1 are unequal to zero, the expression that results from substituting Eqs. (12) and (13) into Eqs. (10) and (11) no longer has an analytical solution. Substituting Eq. (13) into Eqs. (10) and (11) yields
i1520-0442-23-3-757-e14
Similarly, Eq. (12) can be substituted into Eqs. (10) and (11) to give (see appendix A)
i1520-0442-23-3-757-e15
Combining Eqs. (14) and (15) and defining a thermal Rossby number,
i1520-0442-23-3-757-e16
one obtains
i1520-0442-23-3-757-e17
When the heating profile is symmetric, Eq. (17) becomes
i1520-0442-23-3-757-e18
from which the HH80 scaling relation,
i1520-0442-23-3-757-e19
is recovered. When ϕ0 and ϕ1 are unequal to zero, an analytical solution for Eq. (17) is impossible.

3. A perturbation analysis for small asymmetric heating profiles

Here, the case with ϕ0 < 0 and ϕ1 < 0 will be discussed, but this choice is not essential for the derivation. To retain the largest amount of generality we write
i1520-0442-23-3-757-e20
i1520-0442-23-3-757-e21
Together with Eq. (17), a system of two equations is obtained with four small parameters: three unknowns (ϵS, ϵW, and ϕ1) and one forcing parameter ϕ0. It is evident that another equation is needed, but first the expressions for ϕS and ϕW are substituted in Eq. (17). At first order, only terms that are linear in the small parameters are retained. Equation (17) becomes, at first order,
i1520-0442-23-3-757-e22
Substituting the expressions for ϕW and ϕS gives
i1520-0442-23-3-757-e23
i1520-0442-23-3-757-e24
It is evident that ϵS = −ϵW. We write ϵW = −ϵS = ϵ. The final relation between the small parameters becomes
i1520-0442-23-3-757-e25
To solve Eq. (25) an extra equation is needed, which comes from demanding continuity of θ in ϕ1. Equation (14) expresses θ(ϕ1) − θf (ϕ1) as a function of ϕW, ϕ1, and ϕ0. However, the exact same relation, with ϕW replaced by ϕS, must hold when the SH cell is considered. As a result, all terms containing ϕW in Eq. (14) must equal the same expression, with ϕW replaced by ϕS. We obtain
i1520-0442-23-3-757-e26
When expanding ϕW and ϕS in ϵ, in the first-order balance one finds
i1520-0442-23-3-757-e27
Substituting this into Eq. (25) gives
i1520-0442-23-3-757-e28

The dividing latitude between the Northern Hemisphere (NH) and SH Hadley cells shifts with the latitude of maximum heating, and in the linear regime this shift is 6 times larger than the shift of the latitude of maximum heating.

4. The second-order expansion

a. Linear in ϕH but nonlinear in ϕ0

The perturbation method can be continued with progressively higher-order expansions for ϕW and ϕS, but care has to be taken to remain consistent with the small-angle approximation. A second-order expansion, however, is still useful. Figure 4 of LH88 shows that ϕW and ϕS display significant curvature as functions of ϕ0. Here, ϕS first decreases as a function of ϕ0, consistent with the first-order linear expansion. However, for larger ϕ0, ϕS increases as a function of ϕ0. Also, the initial linear gradient of these curves is corrected by higher-order expansions in ϕH. Here, the second-order expansion in ϕ0 is considered without relaxing the (linear) small-angle approximation. This expansion is not fully consistent for small ϕ0, but it facilitates the derivation of the full second-order problem. In the full second-order problem, a nonlinear small-angle approximation is needed, which allows a higher-order expansion in ϕH. This affects the linear relation between ϕ0 and all other variables.

When using a linear small-angle approximation, the form of ϕW and ϕS is anticipated to become to second order, as
i1520-0442-23-3-757-e29
i1520-0442-23-3-757-e30
The second-order approximation of Eq. (17) is
i1520-0442-23-3-757-e31
and a similar equation for ϕS holds. Substituting the expressions for ϕW (ϕS) and retaining only those terms that are second order in the small parameters ϵ, δ, ϕ1, and ϕ0 gives an expression for ϕW and ϕS that can be solved directly as
i1520-0442-23-3-757-e32
A quadratic term could be added to the equation for ϕ1, but it turns out that this term must be zero. Thus,
i1520-0442-23-3-757-e33
i1520-0442-23-3-757-e34
i1520-0442-23-3-757-e35

b. Nonlinear in ϕH and ϕ0

To be fully consistent, the sine and cosine terms that were linearized in the small-angle approximation have to be expanded to higher order too. It must be recalled that ϕW is (much) larger than either ϕ1 or ϕ0, and that we only need to retain terms that are second order in ϕ0 and ϕ1, or larger. When Eqs. (14) and (15) are expanded and combined together (see appendix B), one obtains
i1520-0442-23-3-757-e36
In section 4a, it was found that the first-order expansion of ϕW and ϕS in ϕ0 only contains antisymmetric terms, while the second-order expansion only contains symmetric terms. Equation (36), however, allows for symmetric and antisymmetric terms at both orders in ϕ0. We have to rewrite this as
i1520-0442-23-3-757-e37
i1520-0442-23-3-757-e38
i1520-0442-23-3-757-e39
The symmetric part of the first-order expansion of Eq. (36) becomes
i1520-0442-23-3-757-e40
which gives
i1520-0442-23-3-757-e41
Having β = O(ϕH3), we demand that ϕ0, ϵ, and ϕ1 are O(ϕH3). In reality, this only holds for a certain range of ϕ0, but it facilitates the derivation when it is assumed that β and ϵ are each of the same order. In practice, relaxing this demand does not change the solution in a fundamental way. It only determines the amount of terms with ϕH that have to be retained, relative to the amount of terms with ϕ0. Because we are only interested in a first-order nonlinear correction to the linear results of the previous paragraph, it is convenient to take the first-order symmetric and antisymmetric terms to be of equal size.
The symmetric part of the second-order expansion of Eq. (36) is (see appendix C)
i1520-0442-23-3-757-e42
Comparing Eqs. (42) and (32) we see that, like in case of β, curvature terms enter the equation for δ2, which depend on higher-order terms in ϕH. The antisymmetric terms in the second-order expansion of Eq. (36) give rise to the relation (appendix C)
i1520-0442-23-3-757-e43
From the full second-order expansion of Eq. (26), one obtains (appendix C)
i1520-0442-23-3-757-e44
Taking Eqs. (43) and (44) together then gives
i1520-0442-23-3-757-e45
The final relations for ϕW, ϕS, and ϕ1 become
i1520-0442-23-3-757-e46
i1520-0442-23-3-757-e47
i1520-0442-23-3-757-e48

These relations are displayed in Fig. 3, which can be compared to Fig. 4 in LH88. For small ϕ0 the agreement is good. The most remarkable difference is the behavior of ϕ1, which becomes smaller for larger ϕH, contrary to the functional relationship displayed in LH88. This discrepancy disappears at higher order. Then, terms of the form ϕ02/ϕH and ϕ03/ϕH4, etc., are added to Eqs. (46)(48). They bend the curve of ϕ1 for smaller ϕH below the curve for larger ϕH.

In addition, the width of the solstitial winter cell Ww can be compared to the width of the equinoctial cell Weq,
i1520-0442-23-3-757-e49
The width of the summer cell has the same form, but a minus term between the brackets instead of this positive term.

5. Numerical solutions from a coupled climate model

In this section the analytical solutions are compared to results from the CCM simulation: the ECHAM5/Max Planck Institute Ocean Model (MPI-OM) climate model (Marsland et al. 2003; Roeckner et al. 2003). With this model, a 17-member ensemble of model simulations was performed over the 1950–2100 period, with increasing greenhouse gas concentrations, as part of the ESSENCE project (Sterl et al. 2008), plus some additional dedicated ensemble experiments. Here, two 5-member ensemble simulations are compared. In one ensemble, a freshwater anomaly of 1 Sv (1 Sv ≡ 106 m3 s−1) was uniformly applied over the northern North Atlantic between 50° and 70°N from 2001 onward, starting from five 1 January 2001 states of the 17-member ensemble baseline experiment. The 5-member subset of the 17-member baseline experiment is the reference ensemble. After this date, the two ensembles develop differently. In the hosing ensemble, the additional freshwater supply leads to a collapse of the THC within 20 yr. After this collapse, the ensemble mean difference between the two ensembles hardly evolves. An almost stationary anomaly pattern prevails that can be seen as the fingerprint of the THC collapse on the atmosphere (Laurian et al. 2010). At first order, this fingerprint is independent of the precise mean atmospheric state (climate). This is probably because of the fact that the THC weakens only moderately during the twenty-first century in the baseline experiment. In this study, the atmospheric fingerprint of the THC collapse is defined as the difference in ensemble mean between the two 5-member ensembles over the 2091–2100 period.

a. The shift in rising branch

First, the relation
i1520-0442-23-3-757-e50
is examined. We take the annual mean circulation as a base state. Figure 4 shows the zonally averaged SST anomaly associated with a THC collapse, together with a linear regression of the temperature profile to latitude. Here, ΔTC is estimated to be 1.466°C. As in HH80, ΔH is taken to be the temperature difference on earth when forced with annual mean solar fluxes, namely, 100°C. Then, ϕ0 is 0.838° (noting that 0.8° equals 0.014 radians). Furthermore, ϕH = 32.5° = 0.569 radians. Substituting these values into Eq. (50) it follows that ϕ1 is 4.4°. Without the nonlinear term the number would be 5.0°; the nonlinear term adds a correction of about 10%. Figure 5 shows the displacement of the dividing line between the annual mean SH and NH Hadley cells as a function of height in the CCM. Here, ϕ1 varies between 1° and 8°, depending on the latitude, with lower values near the surface and in the stratosphere and a maximum near 400 hPa. The average value of ϕ1 is 4.2°, the difference between the CCM estimate and the theoretical value is less than 10%. This is also true when the seasonally varying circulation is taken as a base state. The agreement, however, is somewhat fortuitous, given the large variation that ϕ1 displays with height. Also, Eq. (50) can be applied to the seasonal cycle. For the June–September (JJAS) minus December–March (DJFM) circulation, the theory predicts 15.1° versus 14.3° observed, and for the more extreme July minus January circulation the prediction is 17.5° versus 15.3° observed. The larger numbers predicted by theory are consistent with the fact that the relation between ϕ1 and ϕ0 is a downward-bending curve for larger ϕ0 (see Fig. 4 of LH88).

The width of the winter and summer cells can be estimated from Eq. (49). The difference between the two is for a large part determined by ϕ1. Equation (49), however, significantly overestimates the difference in width because the poleward extent of the Hadley cells is not correctly captured. This issue is further discussed in section 5c.

b. The strength of the anomalous meridional circulation

The strength of the anomalous cross-equatorial cell that results from a THC collapse is 3.03 1010 kg s−1 (Fig. 6). This can be compared to the solstitial circulation that features a cross-equatorial cell of 24.54 1010 kg s−1. The strength of the solstitial circulation is obtained by subtracting the time-averaged January and July circulations from the annually averaged meridional mass transport. The annually mean meridional circulation is asymmetric (Fig. 6), but the yearly average of the NH and SH Hadley cells is 9.43 1010 kg s−1.

The HH80 and LH88 models for the Hadley circulation systematically underpredict the strength of the Hadley circulation because of a too-broad heating pattern (Hou and Lindzen 1992). Also, the absolute magnitude of the Hadley cell cannot be assessed without accounting for the role of eddies (Walker and Schneider 2005, 2006). However, if the rectification resulting from eddies scales with the strength of the mean flow, the scaling for the strength of the flow still holds in a relative sense. Therefore, only its relative strength will be compared with the scaling that results from the expansion discussed before. The starting point for estimating the mass flux in the Hadley cell is the notion that the strength of the Hadley circulation is proportional to the difference between θ and θe (LH88). The relevant scaling for the symmetric case is given by Eq. (18) from HH80 as
i1520-0442-23-3-757-e51
With the aid of Eqs. (10) and (11), this equation is reformulated for the asymmetric case:
i1520-0442-23-3-757-e52
Here, Ψas is the equatorially asymmetric overturning circulation driven by off-equatorially heating. To arrive at an equation that solely depends on the forcing parameters, the functional dependence of the temperature on ϕ has to be substituted. This is possible with the use of Eq. (14), which describes the temperature deviation from the radiative equilibrium temperature. Substituting Eq. (14) into Eq. (52), and using the expansion given by Eqs. (46)(48), one finds that for small ϕ (see appendix D),
i1520-0442-23-3-757-e53
For ϕ0 = 0 and ϕ1 = 0 this equation is equal to Eq. (18) of HH80, apart from the fourth-order terms that were neglected here. The part within brackets in Eq. (53) is expressed in ϕ0, while the integrand is expressed in terms of ϕ1. The reason for this is that the integral bounds are exact, while the formula within brackets is based on an expansion in ϕ0. Replacing ϕ0 by ϕ1, here, means that higher-order terms are introduced. These terms are significant for the solstitial circulation (for which this expansion strictly spoken is not valid), but terms of similar order were neglected in that expansion. As a result, the equation becomes unbalanced when ϕ0 is replaced by ϕ1.
The anomalous overturning circulation can easily be derived from Eq. (53). This circulation peaks at the equator, where the mean symmetric circulation is zero. For ϕ = 0, the anomalous cross-equatorial cell equals the asymmetric solution,
i1520-0442-23-3-757-e54

In the CCM ensemble, the ratio between the cross-equatorial cell associated with the seasonal cycle and the cell resulting from a THC collapse is 24.54/3.03 = 8.1. For the THC collapse ϕ1 = −4.2°, ϕH = 32.5°, and ϕ0 = −0.84°. For the solstitial circulation the values are ϕ1 = −15.3°, ϕH = 30.3°, and ϕ0 = −3.3° (the cross-equatorial SST gradient was estimated from a linear regression to latitude as 5.8°C). These values imply a ratio of 7.8 for the solstitial cell compared to the THC-induced cell. This number compares to the ratio found in the CCM.

c. The poleward extent of the Hadley cell and the strength of the zonal jet

Figure 6 shows that the NH meridional cell increases in strength, but that its outer boundary recedes, while the SH meridional cell weakens and its outer boundary expands after a THC collapse. This behavior is further illustrated by Fig. 7. It is seen that the SH cell slightly expands its outer edge with 0.1°, while the NH cell outer edge shrinks with 0.6°. However, according to Eq. (48), the outer edge of the SH cell should shrink in response to the displacement of maximum heating latitude, whereas the poleward extent of the NH cell should increase. The change in the poleward extent of the Hadley cells, as predicted from theory, is contrary to what the CCM shows.

The reason for the discrepancy is that the poleward extent of the Hadley cells is controlled by baroclinic instability. Angular momentum conservation only applies as long as the vertical shears do not exceed a critical value (Held 2000). If baroclinic instability occurs at a latitude that is smaller than the latitude determined by angular momentum conservation, baroclinic instability determines the poleward extent of the Hadley cell. Lu et al. (2007) found that the Hadley cell expanded and weakened under global warming in a series of Intergovernmental Panel on Climate Change Fourth Assessment Report (IPCC AR4) models, following the Held (2000) scaling, whereas angular momentum conservation predicts the opposite behavior. This result implies that baroclinic instability indeed controls the outer boundary of the Hadley circulation in the present climate. Both in Held (2000) and in Lu et al. (2007), it is assumed that the velocity profile itself does not change, and that the change in the poleward extent of the Hadley cell is determined by changes in tropopause height and static stability. In the case of a THC collapse, this assumption does not apply. The latitudinal temperature profile changes significantly and as a result the balanced zonal wind changes (see Fig. 8). The Held (2000) scaling can still be taken as a starting point to explain this behavior. The flow becomes unstable when
i1520-0442-23-3-757-e55
where β = ∂f /∂y. In Held (2000), it is assumed that when the angular momentum–conserving wind exceeds a critical shear, the Hadley cell is terminated. However, near the Hadley cell terminus, upper-tropospheric flows deviate substantially from angular momentum–conserving winds (Walker and Schneider 2006). Instead, we simply demand thermal wind balance,
i1520-0442-23-3-757-e56
To equate the Eqs. (55) and (56) results in a supercriticality criterion:
i1520-0442-23-3-757-e57
This criterion is almost the same as the one advocated by Korty and Schneider (2008), only here the meridional gradient of the vertically averaged potential temperature is used instead of the surface potential temperature. We arrive at the following scaling,
i1520-0442-23-3-757-e58
The poleward Hadley cell extent is inversely proportional to the isentropic slopes, when the slopes increase the Hadley cell contracts and vice versa.

To calculate the change resulting from a THC collapse, temperature gradients were averaged over the latitude band of 30°–40°, which includes the zero contour of the meridional streamfunction and the subtropical jet velocity maximum. First, the fractional change in ΔVθ0 is considered. With surface warming in the SH and cooling in the NH, the static stability in the SH near the outer boundary of the Hadley cell decreases by 1.2% and it increases in the NH by 2.5%. The fractional change in ∂θ/∂ϕ gives a decrease of 3.2% in the SH and an increase of 6.4% in the NH. The net result is that the SH Hadley cell should expand with 2.0% ≃ 0.6° and the NH Hadley cell should shrink with 3.9% ≃ 1.2°. When comparing with Fig. 7, it is seen that the predicted changes are overestimated. In the CCM, the poleward expansion in the SH is only 0.1°, and the contraction of the outer edge of the NH cell is 0.6°. The Korty and Schneider (2008) criterion overestimates the changes in Hadley cell extent even more. Qualitatively, both criteria predict the change in the Hadley cell extent, but quantitatively there are discrepancies. These are probably related to the neglect of changes in moisture.

There is no simple theory to predict the observed changes in ∂θ/∂ϕ. However, the axisymmetric theory demands that poleward of the Hadley cell terminus θ = θf . This is not strictly true, because eddies transport heat when the meridional flow is absent, but it is still worthwhile to investigate whether the changes in forcing temperature as predicted by theory are consistent with the observed changes in actual temperature. The derivative of Eq. (12) is
i1520-0442-23-3-757-e59
In other words,
i1520-0442-23-3-757-e60
It follows that ∂θf /∂ϕ increases by 2.7% in the NH; it decreases by 2.5% in the SH because of the shift in maximum heating latitude.

Next, the change resulting from a change in absorbed radiation is estimated. The THC collapse induces a large albedo response in the NH, and the signal in the SH is weak and insignificant. In the NH the decrease in absorbed shortwave radiation is 7.2%. This implies a temperature drop of 0.25 × 7.2% = 1.8%, which is equivalent to 3.1°C near the North Pole. This means that ΔH also increases by 3.1%.

According to the analytical model, the decrease in ∂θf /∂ϕ in the SH is 2.5%, compared to a 3.2% decrease of actual temperature gradient in the CCM. The increase in the forcing temperature gradient in the NH is predicted to be 5.8%, compared to a 6.4% increase of the actual temperature gradient in the CCM. The analytically predicted changes in the forcing temperature are consistent with the observed changes in the actual temperature gradient near the outer boundary of the Hadley cells. The heat transport by the eddies is significant, but it does not break up the relation that exists near the terminus of the Hadley cells between the actual temperature changes and changes in the forcing temperature.

6. Conclusions and discussion

Our main conclusions are as follows:

  1. The cross-hemispheric SST anomaly associated with a THC collapse is equivalent to an equatorward shift of the latitude of maximum heating. The dividing latitude between the NH and SH Hadley cells features an equatorward and southward shift that is roughly 5 times larger than the shift in latitude of maximum heating.

  2. The equatorward shift of the mean Hadley circulation resulting from a THC collapse makes the Hadley circulation more symmetric. The anomalous cross-hemispheric circulation features enhanced energy transport from the SH to the NH (the total Hadley circulation features reduced energy transport from the NH to the SH). This cell can be expressed as a polynomial of the shift in heating latitude.

  3. In the NH, the Hadley cell contracts, while the zonal velocities increase and the subtropical jet shifts equatorward. In the SH the opposite occurs. This behavior is explained by assuming that the outer boundary of the Hadley cell is determined by baroclinic instability. Then, its poleward extent scales with the inverse of the isentropic slope at subtropical latitudes.

The qualitative agreement between the scaling laws derived from the dry axisymmetric model of the Hadley circulation and the CCM results may come as a surprise. The role of the oceans in the Hadley circulation primarily is changing the thermal forcing by (moist) convection. Also, the altered thermal forcing has a distinct zonal pattern related to the land–sea distribution (Clement 2006). This zonal asymmetry is even further strengthened by the role of the Atlantic as the primary forcing agent when the THC collapses, whereas the other oceans play a smaller role. Therefore, the changes in atmospheric moist content are large and show a distinct zonal pattern.

Also, eddies may compromise the axisymmetric model. The zonal flow in the Hadley cells deviates substantially from the angular momentum–conserving flow because eddy momentum fluxes are large (Walker and Schneider 2006), especially near the terminus of the Hadley cell. As a result, they significantly alter the momentum balance. Indeed, the CCM deviates qualitatively from the scalings that we derived from the HH80 and LH88 theories, with regard to the poleward extent of the Hadley cell and changes in zonal velocity. Both are governed by baroclinic instability (Held 2000). When this aspect is included in the simple axisymmetric model, the CCM response is consistent with the appropriate scaling. This also applies for the widening of the Hadley cell resulting from greenhouse warming (e.g., Lu et al. 2007).

The use of the Held (2000) criterion, however, has been criticized by Korty and Schneider (2008). Therefore, Korty and Schneider advocated an alternative criterion that was based on supercriticality, namely, one that measures the depth of the eddy entropy flux. The scaling that we used consists of a modification to the Held (2000) scaling, consistent with the arguments of Korty and Schneider, but with a slightly different outcome.

This deviation from the HH80 and LH88 model stresses the role of eddies in affecting the Hadley circulation. Also, the relation between ENSO and variations in the Hadley cell strength point to an important role of eddies, with ENSO affecting eddy stresses in the tropical momentum balance (Quan et al. 2004; Caballero 2007). The main difference between tropical forcing by ENSO and by a THC collapse is that the THC collapse is associated with an SST anomaly with a strong cross-hemispheric SST gradient, which imposes a similar forcing as seasonally varying insolation; whereas the ENSO-related tropical SST anomaly mainly consists of zonal contrasts, which affect the tropical stationary waves.

The latitude where the summer and winter cell divide shifts southward, and as a result the latitude of the intertropical convergence zone (ITCZ) shifts southward, in response to a THC collapse (see, e.g., Zhang and Delworth 2005). Such a shift also occurs either when large-scale NH SST cooling is applied in an atmosphere–slab ocean model (Broccoli et al. 2006), or when NH land and sea ice cover changes (Chiang and Bitz 2005). With the present theory, it is clear why all of these forcings are equivalent. Instrumental in the southward shift of the ITCZ is the decrease in the latitude of maximal heating ϕ0, which follows from imposing a cross-hemispheric SST anomaly. Such an anomaly immediately causes a response in the tropical Hadley circulation, with a southward shift of the dividing line between the NH and SH cells ϕ1, which is equivalent to a southward shift of the annual mean position of the ITCZ.

This southward shift in latitude of maximum heating dominates the Hadley cell response to reduced ocean heat transport when the THC collapses. Clement (2006) argues that ocean heat transport weakens the symmetric Hadley circulation in favor of the asymmetric solstitial cell. Reduced ocean heat transport resulting from a THC collapse then enhances the symmetric Hadley circulation but weakens the solstitial cell. In the CCM both cells weaken. The effect of reduced ocean heat transport in enhancing the symmetric cell is offset by the equatorward shift of the latitude of maximum heating that acts to decrease the symmetric Hadley cell.

It has often been surmised that reduced ocean heat transport associated with a THC collapse would be compensated by increased atmospheric heat transport by the Bjerknes compensation (see, e.g., Cheng et al. 2007; Vellinga and Wu 2008). There is a tendency for the atmosphere to compensate for reduced ocean heat transport in the CCM, but this compensation is far from complete and a different climate state with a different radiation balance results after the THC collapses. The anomalous meridional mass flux that arises because of this different radiation balance consists of a cross-equatorial cell. This cell is maintained by anomalous TOA net downward radiation in the SH tropics and anomalous TOA net upward radiation in the NH tropics. At midlatitudes and subpolar latitudes, the anomalous net TOA radiation is upward, consistent with reduced northward heat transport in the NH. These net changes in radiation balance are dominated by a large-scale cloud response (Laurian et al. 2010), because clear-sky long- and shortwave radiation changes almost compensate for one another (Vellinga and Wu 2008).

The enhanced northward angular momentum transport by the NH Hadley cell must be associated with enhanced angular momentum transport by the eddy-driven Ferrel cell (Cheng et al. 2007), and as a result an increase in NH transient baroclinic wave activity occurs. The NH subtropical jet intensifies but also shifts southward. This affects the NH eddy-driven jet and storm tracks as well (Lee and Kim 2003). All of these changes point to an intricate link between a THC collapse, a response in the tropical ocean and atmospheric circulation, and the extratropical atmospheric response. Note that for a large part the extratropical atmospheric response is not directly forced by extratropical SST changes, but rather is mediated by the tropical atmospheric response to a tropical SST change, most noticeably the cross-equatorial SST gradient.

Seager and Battisti (2007) discussed possible causes for abrupt climate change and advocated for an active role for the tropics as an alternative for what they called the THC theory. Although they noted that a THC collapse induces changes in the tropics, they concluded that the overall atmospheric response was too small to explain the paleorecord. Their alternative theory was based on a bifurcation in the position and strength of the eddy-driven jets (following Lee and Kim 2003). The results presented in this paper strongly suggest that this competition between the tropics and the THC is artificial. In reality, all mechanisms discussed previously appear to be intrinsically linked. The THC collapse induces tropical SST changes resulting from ocean adjustment and a wind–evaporation–SST (WES) feedback (Chiang and Bitz 2005), and this, in turn, affects the Hadley circulation. The changed Hadley circulation, being associated with a shift in the subtropical jets, impacts midlatitude storm tracks and the eddy-driven jet.

Acknowledgments

Computer resources were funded by the National Computing Facilities Foundation (NCF). We thank Michael Kliphuis for technical support and three anonymous reviewers for their constructive comments.

REFERENCES

  • Broccoli, A. J., K. A. Dahl, and R. J. Stouffer, 2006: Response of the ITCZ to Northern Hemisphere cooling. Geophys. Res. Lett., 33 , L01702. doi:10.1029/2005GL024546.

    • Search Google Scholar
    • Export Citation
  • Caballero, R., 2007: Role of eddies in the interannual variability of Hadley cell strength. Geophys. Res. Lett., 34 , L22705. doi:10.1029/2007GL030971.

    • Search Google Scholar
    • Export Citation
  • Cheng, W., C. M. Bitz, and J. C. H. Chiang, 2007: Adjustment of the global climate to an abrupt slowdown of the Atlantic meridional overturning circulation. Ocean Circulation: Mechanisms and Impacts, Geophys. Monogr., Vol. 173, Amer. Geophys. Union, 295–314.

    • Search Google Scholar
    • Export Citation
  • Chiang, J. C. H., and C. M. Bitz, 2005: Influence of high-latitude ice cover on the marine Intertropical Convergence Zone. Climate Dyn., 25 , 477496.

    • Search Google Scholar
    • Export Citation
  • Clement, A. C., 2006: The role of the ocean in the seasonal cycle of the Hadley circulation. J. Atmos. Sci., 63 , 33513365.

  • Fang, M., and K. T. Tung, 1996: A simple model of nonlinear Hadley circulation with an ITCZ: Analytical and numerical solutions. J. Atmos. Sci., 53 , 12411261.

    • Search Google Scholar
    • Export Citation
  • Fang, M., and K. T. Tung, 1999: Time-dependent nonlinear Hadley circulation. J. Atmos. Sci., 56 , 17971807.

  • Held, I. M., 2000: The general circulation of the atmosphere. Woods Hole Oceanographic Institution Tech Rep. 54 pp. [Available online at http://www.whoi.edu/fileserver.do?id=21464&pt=10&p=17332].

    • Search Google Scholar
    • Export Citation
  • Held, I. M., and A. Y. Hou, 1980: Nonlinear axially symmetric circulations in a nearly inviscid atmosphere. J. Atmos. Sci., 37 , 515534.

    • Search Google Scholar
    • Export Citation
  • Hou, A. Y., and R. S. Lindzen, 1992: The influence of concentrated heating on the Hadley circulation. J. Atmos. Sci., 49 , 12331241.

  • James, I. N., 1994: Introduction to Circulating Atmospheres. Cambridge University Press, 422 pp.

  • Korty, R. L., and T. Schneider, 2008: Extent of Hadley circulations in dry atmospheres. Geophys. Res. Lett., 35 , L23803. doi:10.1029/2008GL035847.

    • Search Google Scholar
    • Export Citation
  • Laurian, A., S. S. Drijfhout, W. Hazeleger, and B. van den Hurk, 2010: Response of the Western European climate to a collapse of the thermohaline circulation. Climate Dyn., doi:10.1007/s00382-008-0513-4, in press.

    • Search Google Scholar
    • Export Citation
  • Lee, S., and H-K. Kim, 2003: The dynamical relationship between subtropical and eddy-driven jets. J. Atmos. Sci., 60 , 14901503.

  • Lindzen, R. S., and A. Y. Hou, 1988: Hadley circulations for zonally averaged heating centered off the equator. J. Atmos. Sci., 45 , 24162427.

    • Search Google Scholar
    • Export Citation
  • Lu, J., G. A. Vecchi, and T. Reichler, 2007: Expansion of the Hadley cell under global warming. Geophys. Res. Lett., 34 , L06805. doi:10.1029/2006GL028443.

    • Search Google Scholar
    • Export Citation
  • Manabe, S., and R. J. Stouffer, 1994: Multiple century response of a coupled ocean–atmosphere model to an increase of atmospheric carbon dioxide. J. Climate, 7 , 523.

    • Search Google Scholar
    • Export Citation
  • Marsland, S. J., H. Haak, J. H. Jungclaus, M. Latif, and F. Röske, 2003: The Max-Planck-Institute global ocean/sea ice model with orthogonal curvilinear coordinates. Ocean Modell., 5 , 91127.

    • Search Google Scholar
    • Export Citation
  • Quan, X-W., H. F. Diaz, and M. P. Hoerling, 2004: Change of the tropical Hadley cell since 1950. The Hadley Cell: Past, Present and Future, H. F. Diaz and R. S. Bradley, Eds., Cambridge University Press, 85–120.

    • Search Google Scholar
    • Export Citation
  • Roeckner, E., and Coauthors, 2003: The atmospheric general circulation model ECHAM 5. Part I: Model description. Max-Planck Institute Rep. 349, 127 pp.

    • Search Google Scholar
    • Export Citation
  • Schneider, E. K., 1977: Axially symmetric steady-state models of the basic state for instability and climate studies. Part II: Nonlinear calculations. J. Atmos. Sci., 34 , 280297.

    • Search Google Scholar
    • Export Citation
  • Seager, R., and D. S. Battisti, 2007: Challenges to our understanding of the general circulation: Abrupt climate change. The Global Circulation of the Atmosphere: Phenomena, Theory, Challenges, T. Schneider and A. H. Sobel, Eds., Princeton University Press, 332–372.

    • Search Google Scholar
    • Export Citation
  • Shaffrey, L., and R. Sutton, 2006: Bjerknes compensation and the decadal variability of energy transports in a coupled climate model. J. Climate, 19 , 11671181.

    • Search Google Scholar
    • Export Citation
  • Sterl, A., and Coauthors, 2008: When can we expect extremely high surface temperatures? Geophys. Res. Lett., 35 , L14703. doi:10.1029/2008GL034071.

    • Search Google Scholar
    • Export Citation
  • van der Schrier, G., S. S. Drijfhout, W. Hazeleger, and L. Noulin, 2007: Increasing the Atlantic subtropical jet cools the circum-North Atlantic region. Meteor. Z., 16 , 675684. doi:10.1127/0941-2948/2007/0252.

    • Search Google Scholar
    • Export Citation
  • van der Swaluw, E., S. S. Drijfhout, and W. Hazeleger, 2007: Bjerknes compensation at high northern latitudes: The ocean forcing the atmosphere. J. Climate, 20 , 60236032.

    • Search Google Scholar
    • Export Citation
  • Vellinga, M., and R. A. Wood, 2002: Global climatic impacts of a collapse of the Atlantic thermohaline circulation. Climatic Change, 54 , 251267.

    • Search Google Scholar
    • Export Citation
  • Vellinga, M., and P. Wu, 2008: Relations between northward ocean and atmosphere energy transports in a coupled climate model. J. Climate, 21 , 561575.

    • Search Google Scholar
    • Export Citation
  • Walker, C. C., and T. Schneider, 2005: Response of idealized Hadley circulations to seasonally varying heating. Geophys. Res. Lett., 32 , L06813. doi:10.1029/2004GL022304.

    • Search Google Scholar
    • Export Citation
  • Walker, C. C., and T. Schneider, 2006: Eddy influences on Hadley circulations: Simulations with an idealized GCM. J. Atmos. Sci., 63 , 33333350.

    • Search Google Scholar
    • Export Citation
  • Zhang, R., and T. L. Delworth, 2005: Simulated tropical response to a substantial weakening of the Atlantic thermohaline circulation. J. Climate, 18 , 18531860.

    • Search Google Scholar
    • Export Citation

APPENDIX A

Energy Conservation

Substituting Eq. (12) into Eqs. (10) and (11) gives
i1520-0442-23-3-757-ea1
The LHS of this equation can be written as
i1520-0442-23-3-757-ea2
The first term on the RHS becomes
i1520-0442-23-3-757-ea3
the second term on the RHS is
i1520-0442-23-3-757-ea4
and the last term on the RHS is
i1520-0442-23-3-757-ea5

APPENDIX B

Equation for the Poleward Boundary

When the small-angle approximation in Eq. (14) is expanded further, one obtains
i1520-0442-23-3-757-eb1
Also, we find
i1520-0442-23-3-757-eb2
and
i1520-0442-23-3-757-eb3
Thus, the next-order expansion in the small angle of Eq. (14) gives
i1520-0442-23-3-757-eb4
In a similar way, Eq. (15) is modified. The LHS of this equation can be written as
i1520-0442-23-3-757-eb5
The first term on the RHS becomes
i1520-0442-23-3-757-eb6
the second term on the RHS is
i1520-0442-23-3-757-eb7
and the last term on the RHS is
i1520-0442-23-3-757-eb8
The next-order expansion in small angle of Eq. (15) then gives
i1520-0442-23-3-757-eb9
Combining Eqs. (B4) and (B9) then gives Eq. (36).

APPENDIX C

Perturbation in Second Order

The symmetric terms in the second-order expansion of Eq. (36) become
i1520-0442-23-3-757-ec1
This equation can be written as
i1520-0442-23-3-757-ec2
Substituting for β, ϵ, and ϕ1 gives
i1520-0442-23-3-757-ec3
The final result is approximately
i1520-0442-23-3-757-ec4
The antisymmetric terms in the second-order expansion of Eq. (36) become
i1520-0442-23-3-757-ec5
This gives the relation
i1520-0442-23-3-757-ec6
After substituting for ϵ, β, and ϕ1, this becomes either
i1520-0442-23-3-757-ec7
or
i1520-0442-23-3-757-ec8
The full second-order expansion of Eq. (26) is
i1520-0442-23-3-757-ec9
The antisymmetric terms in this equation give
i1520-0442-23-3-757-ec10
This leads to either the relation
i1520-0442-23-3-757-ec11
or
i1520-0442-23-3-757-ec12
which is equivalent to
i1520-0442-23-3-757-ec13

APPENDIX D

Cross-Equatorial Mass Flux

The starting point is Eq. (52),
i1520-0442-23-3-757-ed1
It is assumed that Ψan peaks at ϕ = 0, so Ψan = Ψas(0). To obtain the response in ϕ1, this equation is developed in second order. For convenience, higher-order expansions in ϕH are neglected because ϕ1 now can be of the same order of magnitude as ϕH. Therefore, the expansion of section 4b is used here. The first term on the RHS is written as
i1520-0442-23-3-757-ed2
The second part is of higher order and can be neglected. The third part is written as
i1520-0442-23-3-757-ed3
and the last part can be written as
i1520-0442-23-3-757-ed4

Fig. 1.
Fig. 1.

The tropical Atlantic 2-m air temperature response associated with a collapse of the THC. The pattern is obtained from the ensemble-averaged mean air temperature over 2090–2100 for five simulations following the Special Report on Emissions Scenarios (SRES) A1b scenario, with an additional freshwater supply of 1 Sv in the northern North Atlantic from 2100 onward, and subtracting this pattern from the 2-m air temperature for five simulations with the same state in year 2100, but without the anomalous freshwater supply after 2100.

Citation: Journal of Climate 23, 3; 10.1175/2009JCLI3159.1

Fig. 2.
Fig. 2.

Schematic of the Hadley circulation with a heating maximum in the SH (boreal winter).

Citation: Journal of Climate 23, 3; 10.1175/2009JCLI3159.1

Fig. 3.
Fig. 3.

Here, ϕ1, ϕW, and ϕS are in radians as functions of ϕ0 (x axis) for ΔH = ⅓ (solid lines) and ΔH = ⅙ (dashed lines), being equivalent to 97° and 48°C, respectively. These values have been chosen to coincide with those used by LH88 in their Fig. 4.

Citation: Journal of Climate 23, 3; 10.1175/2009JCLI3159.1

Fig. 4.
Fig. 4.

The difference in the zonally averaged tropical SST that is caused by a collapse of the THC (dashed line) and the linear regression against the latitude of this temperature profile (solid line).

Citation: Journal of Climate 23, 3; 10.1175/2009JCLI3159.1

Fig. 5.
Fig. 5.

The displacement of the dividing line between the SH and NH Hadley cells as a function of height (dashed line), resulting from a THC collapse. The star symbol denotes the theoretical value.

Citation: Journal of Climate 23, 3; 10.1175/2009JCLI3159.1

Fig. 6.
Fig. 6.

The ensemble-averaged meridional overturning difference associated with a THC collapse as obtained in ESSENCE, overplotted with contours of the annual mean overturning in the control climate (1010 kg s−1).

Citation: Journal of Climate 23, 3; 10.1175/2009JCLI3159.1

Fig. 7.
Fig. 7.

The displacement of the outer boundaries of the SH (solid line) and NH (dashed line) Hadley cells as a function of height, resulting from a THC collapse.

Citation: Journal of Climate 23, 3; 10.1175/2009JCLI3159.1

Fig. 8.
Fig. 8.

The difference in the zonally averaged zonal velocity resulting from a collapse of the THC (m s−1). Contours denote mean values in the control climate.

Citation: Journal of Climate 23, 3; 10.1175/2009JCLI3159.1

Save
  • Broccoli, A. J., K. A. Dahl, and R. J. Stouffer, 2006: Response of the ITCZ to Northern Hemisphere cooling. Geophys. Res. Lett., 33 , L01702. doi:10.1029/2005GL024546.

    • Search Google Scholar
    • Export Citation
  • Caballero, R., 2007: Role of eddies in the interannual variability of Hadley cell strength. Geophys. Res. Lett., 34 , L22705. doi:10.1029/2007GL030971.

    • Search Google Scholar
    • Export Citation
  • Cheng, W., C. M. Bitz, and J. C. H. Chiang, 2007: Adjustment of the global climate to an abrupt slowdown of the Atlantic meridional overturning circulation. Ocean Circulation: Mechanisms and Impacts, Geophys. Monogr., Vol. 173, Amer. Geophys. Union, 295–314.

    • Search Google Scholar
    • Export Citation
  • Chiang, J. C. H., and C. M. Bitz, 2005: Influence of high-latitude ice cover on the marine Intertropical Convergence Zone. Climate Dyn., 25 , 477496.

    • Search Google Scholar
    • Export Citation
  • Clement, A. C., 2006: The role of the ocean in the seasonal cycle of the Hadley circulation. J. Atmos. Sci., 63 , 33513365.

  • Fang, M., and K. T. Tung, 1996: A simple model of nonlinear Hadley circulation with an ITCZ: Analytical and numerical solutions. J. Atmos. Sci., 53 , 12411261.

    • Search Google Scholar
    • Export Citation
  • Fang, M., and K. T. Tung, 1999: Time-dependent nonlinear Hadley circulation. J. Atmos. Sci., 56 , 17971807.

  • Held, I. M., 2000: The general circulation of the atmosphere. Woods Hole Oceanographic Institution Tech Rep. 54 pp. [Available online at http://www.whoi.edu/fileserver.do?id=21464&pt=10&p=17332].

    • Search Google Scholar
    • Export Citation
  • Held, I. M., and A. Y. Hou, 1980: Nonlinear axially symmetric circulations in a nearly inviscid atmosphere. J. Atmos. Sci., 37 , 515534.

    • Search Google Scholar
    • Export Citation
  • Hou, A. Y., and R. S. Lindzen, 1992: The influence of concentrated heating on the Hadley circulation. J. Atmos. Sci., 49 , 12331241.

  • James, I. N., 1994: Introduction to Circulating Atmospheres. Cambridge University Press, 422 pp.

  • Korty, R. L., and T. Schneider, 2008: Extent of Hadley circulations in dry atmospheres. Geophys. Res. Lett., 35 , L23803. doi:10.1029/2008GL035847.

    • Search Google Scholar
    • Export Citation
  • Laurian, A., S. S. Drijfhout, W. Hazeleger, and B. van den Hurk, 2010: Response of the Western European climate to a collapse of the thermohaline circulation. Climate Dyn., doi:10.1007/s00382-008-0513-4, in press.

    • Search Google Scholar
    • Export Citation
  • Lee, S., and H-K. Kim, 2003: The dynamical relationship between subtropical and eddy-driven jets. J. Atmos. Sci., 60 , 14901503.

  • Lindzen, R. S., and A. Y. Hou, 1988: Hadley circulations for zonally averaged heating centered off the equator. J. Atmos. Sci., 45 , 24162427.

    • Search Google Scholar
    • Export Citation
  • Lu, J., G. A. Vecchi, and T. Reichler, 2007: Expansion of the Hadley cell under global warming. Geophys. Res. Lett., 34 , L06805. doi:10.1029/2006GL028443.

    • Search Google Scholar
    • Export Citation
  • Manabe, S., and R. J. Stouffer, 1994: Multiple century response of a coupled ocean–atmosphere model to an increase of atmospheric carbon dioxide. J. Climate, 7 , 523.

    • Search Google Scholar
    • Export Citation
  • Marsland, S. J., H. Haak, J. H. Jungclaus, M. Latif, and F. Röske, 2003: The Max-Planck-Institute global ocean/sea ice model with orthogonal curvilinear coordinates. Ocean Modell., 5 , 91127.

    • Search Google Scholar
    • Export Citation
  • Quan, X-W., H. F. Diaz, and M. P. Hoerling, 2004: Change of the tropical Hadley cell since 1950. The Hadley Cell: Past, Present and Future, H. F. Diaz and R. S. Bradley, Eds., Cambridge University Press, 85–120.

    • Search Google Scholar
    • Export Citation
  • Roeckner, E., and Coauthors, 2003: The atmospheric general circulation model ECHAM 5. Part I: Model description. Max-Planck Institute Rep. 349, 127 pp.

    • Search Google Scholar
    • Export Citation
  • Schneider, E. K., 1977: Axially symmetric steady-state models of the basic state for instability and climate studies. Part II: Nonlinear calculations. J. Atmos. Sci., 34 , 280297.

    • Search Google Scholar
    • Export Citation
  • Seager, R., and D. S. Battisti, 2007: Challenges to our understanding of the general circulation: Abrupt climate change. The Global Circulation of the Atmosphere: Phenomena, Theory, Challenges, T. Schneider and A. H. Sobel, Eds., Princeton University Press, 332–372.

    • Search Google Scholar
    • Export Citation
  • Shaffrey, L., and R. Sutton, 2006: Bjerknes compensation and the decadal variability of energy transports in a coupled climate model. J. Climate, 19 , 11671181.

    • Search Google Scholar
    • Export Citation
  • Sterl, A., and Coauthors, 2008: When can we expect extremely high surface temperatures? Geophys. Res. Lett., 35 , L14703. doi:10.1029/2008GL034071.

    • Search Google Scholar
    • Export Citation
  • van der Schrier, G., S. S. Drijfhout, W. Hazeleger, and L. Noulin, 2007: Increasing the Atlantic subtropical jet cools the circum-North Atlantic region. Meteor. Z., 16 , 675684. doi:10.1127/0941-2948/2007/0252.

    • Search Google Scholar
    • Export Citation
  • van der Swaluw, E., S. S. Drijfhout, and W. Hazeleger, 2007: Bjerknes compensation at high northern latitudes: The ocean forcing the atmosphere. J. Climate, 20 , 60236032.

    • Search Google Scholar
    • Export Citation
  • Vellinga, M., and R. A. Wood, 2002: Global climatic impacts of a collapse of the Atlantic thermohaline circulation. Climatic Change, 54 , 251267.

    • Search Google Scholar
    • Export Citation
  • Vellinga, M., and P. Wu, 2008: Relations between northward ocean and atmosphere energy transports in a coupled climate model. J. Climate, 21 , 561575.

    • Search Google Scholar
    • Export Citation
  • Walker, C. C., and T. Schneider, 2005: Response of idealized Hadley circulations to seasonally varying heating. Geophys. Res. Lett., 32 , L06813. doi:10.1029/2004GL022304.

    • Search Google Scholar
    • Export Citation
  • Walker, C. C., and T. Schneider, 2006: Eddy influences on Hadley circulations: Simulations with an idealized GCM. J. Atmos. Sci., 63 , 33333350.

    • Search Google Scholar
    • Export Citation
  • Zhang, R., and T. L. Delworth, 2005: Simulated tropical response to a substantial weakening of the Atlantic thermohaline circulation. J. Climate, 18 , 18531860.

    • Search Google Scholar
    • Export Citation
  • Fig. 1.

    The tropical Atlantic 2-m air temperature response associated with a collapse of the THC. The pattern is obtained from the ensemble-averaged mean air temperature over 2090–2100 for five simulations following the Special Report on Emissions Scenarios (SRES) A1b scenario, with an additional freshwater supply of 1 Sv in the northern North Atlantic from 2100 onward, and subtracting this pattern from the 2-m air temperature for five simulations with the same state in year 2100, but without the anomalous freshwater supply after 2100.

  • Fig. 2.

    Schematic of the Hadley circulation with a heating maximum in the SH (boreal winter).

  • Fig. 3.

    Here, ϕ1, ϕW, and ϕS are in radians as functions of ϕ0 (x axis) for ΔH = ⅓ (solid lines) and ΔH = ⅙ (dashed lines), being equivalent to 97° and 48°C, respectively. These values have been chosen to coincide with those used by LH88 in their Fig. 4.

  • Fig. 4.

    The difference in the zonally averaged tropical SST that is caused by a collapse of the THC (dashed line) and the linear regression against the latitude of this temperature profile (solid line).

  • Fig. 5.

    The displacement of the dividing line between the SH and NH Hadley cells as a function of height (dashed line), resulting from a THC collapse. The star symbol denotes the theoretical value.

  • Fig. 6.

    The ensemble-averaged meridional overturning difference associated with a THC collapse as obtained in ESSENCE, overplotted with contours of the annual mean overturning in the control climate (1010 kg s−1).

  • Fig. 7.

    The displacement of the outer boundaries of the SH (solid line) and NH (dashed line) Hadley cells as a function of height, resulting from a THC collapse.

  • Fig. 8.

    The difference in the zonally averaged zonal velocity resulting from a collapse of the THC (m s−1). Contours denote mean values in the control climate.

All Time Past Year Past 30 Days
Abstract Views 0 0 0
Full Text Views 1816 1271 41
PDF Downloads 171 53 1