revisiting the relationship between jet position, · 1 revisiting the relationship between jet...

23
GEOPHYSICAL RESEARCH LETTERS, VOL. ???, XXXX, DOI:10.1002/, Revisiting the relationship between jet position, 1 forced response and annular mode variability in the 2 southern mid-latitudes 3 Isla R Simpson 1,2 and Lorenzo M Polvani, 2,3 Corresponding author: Isla Simpson, Climate and Global Dynamics Laboratory, National Center for Atmospheric Research, Boulder, Colorado, USA. ([email protected]) 1 Climate and Global Dynamics Laboratory, National Center for Atmospheric Research, Boulder, Colorado, USA 2 Division of Ocean and Climate Physics, Lamont-Doherty Earth Observatory, Palisades, New York USA 3 Department of Applied Physics and Applied Mathematics and Department of Earth and Environmental Sciencies, Columbia University, New York, USA DRAFT February 25, 2016, 10:00am DRAFT

Upload: dinhnhan

Post on 28-Aug-2018

217 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Revisiting the relationship between jet position, · 1 Revisiting the relationship between jet position, 2 forced response and annular mode variability in the 3 southern mid-latitudes

GEOPHYSICAL RESEARCH LETTERS, VOL. ???, XXXX, DOI:10.1002/,

Revisiting the relationship between jet position,1

forced response and annular mode variability in the2

southern mid-latitudes3

Isla R Simpson1,2

and Lorenzo M Polvani,2,3

Corresponding author: Isla Simpson, Climate and Global Dynamics Laboratory, National

Center for Atmospheric Research, Boulder, Colorado, USA. ([email protected])

1Climate and Global Dynamics

Laboratory, National Center for

Atmospheric Research, Boulder, Colorado,

USA

2Division of Ocean and Climate Physics,

Lamont-Doherty Earth Observatory,

Palisades, New York USA

3Department of Applied Physics and

Applied Mathematics and Department of

Earth and Environmental Sciencies,

Columbia University, New York, USA

D R A F T February 25, 2016, 10:00am D R A F T

Page 2: Revisiting the relationship between jet position, · 1 Revisiting the relationship between jet position, 2 forced response and annular mode variability in the 3 southern mid-latitudes

X - 2 SIMPSON & POLVANI: SH JET LATITUDES, JET SHIFTS AND THE SAM

Key Points.

◦ The SH jet position/jet shift correlation is

highly seasonal in CMIP5 models

◦ It is strong in JJA but weak in DJF, oppo-

site to the spread in SAM timescales

◦ Proposed theories for the jet position/jet

shift relationship appear inadequate to ex-

plain this seasonality

Climate models exhibit a wide range in lat-4

itudinal position of the Southern Hemisphere5

westerly jet. Previous work has demonstrated,6

in the annual mean, that models with lower7

latitude jets, exhibit greater poleward jet shifts8

under climate forcings. It has been argued that9

this behavior is due to stronger eddy/mean10

flow feedbacks in models with lower latitude11

jets, as inferred from the timescale of the South-12

ern Annular Mode (SAM). Here we revisit this13

question with a focus on seasonality. Using a14

larger set of models and forcing scenarios from15

the Coupled Model Intercomparison Project,16

phase 5, we find that the jet position/jet shift17

relationship is strong in winter but insignif-18

icant in summer, whereas the model spread19

D R A F T February 25, 2016, 10:00am D R A F T

Page 3: Revisiting the relationship between jet position, · 1 Revisiting the relationship between jet position, 2 forced response and annular mode variability in the 3 southern mid-latitudes

SIMPSON & POLVANI: SH JET LATITUDES, JET SHIFTS AND THE SAM X - 3

in SAM timescales arises primarily in summer,20

with winter timescales similar across models.21

The results, therefore, question previous in-22

terpretations and motivate an improved un-23

derstanding of the spread in model behavior.24

D R A F T February 25, 2016, 10:00am D R A F T

Page 4: Revisiting the relationship between jet position, · 1 Revisiting the relationship between jet position, 2 forced response and annular mode variability in the 3 southern mid-latitudes

X - 4 SIMPSON & POLVANI: SH JET LATITUDES, JET SHIFTS AND THE SAM

1. Introduction

Accurate model predictions of future changes in the Southern Hemisphere (SH) west-25

erlies are necessary for accurate prediction of SH regional climate, Antarctic sea ice cover26

[Sigmond and Fyfe, 2010; Smith et al., 2012], the ocean circulation and its uptake of heat27

and carbon [e.g., Marshall and Speer , 2012; Gent , 2016]. A long standing barrier to our28

confidence in model predictions of the SH westerlies, however, is the current inability to29

simulate their observed climatology and variability.30

Many climate models place the SH zonal mean westerly jet maximum, equatorward of31

the observations [Fyfe and Saenko, 2006; Swart and Fyfe, 2012; Wilcox et al., 2012; Brace-32

girdle et al., 2013] and studies have suggested that the extent of a model’s equatorward33

jet bias correlates with how much the jet shifts in latitude under climate forcings (Kidston34

and Gerber [2010, KG2010, hereafter], Son et al. [2010], Bracegirdle et al. [2013]). Ulti-35

mately, one would want to alleviate this equatorward bias, but such a correlation might36

also serve as an emergent constraint on the future evolution the climate system [Collins37

et al., 2012]. Any such constraint must, however, be grounded on a solid understanding38

of the physical processes involved.39

Our current understanding of the jet latitude/jet shift relationship is that it arises be-40

cause models with lower latitude jets, exhibit stronger eddy/mean flow feedbacks. This41

has been inferred from a relationship between jet latitude and the persistence of the SAM,42

as measured by its decorrelation timescale. Lower latitude jets exhibit greater SAM per-43

sistence, as shown by KG2010 with global coupled models, and several idealized modeling44

studies [Son and Lee, 2005; Gerber and Vallis , 2007; Barnes et al., 2010; Simpson et al.,45

D R A F T February 25, 2016, 10:00am D R A F T

Page 5: Revisiting the relationship between jet position, · 1 Revisiting the relationship between jet position, 2 forced response and annular mode variability in the 3 southern mid-latitudes

SIMPSON & POLVANI: SH JET LATITUDES, JET SHIFTS AND THE SAM X - 5

2010]. The greater persistence of this dominant mode of natural variability is taken to46

indicate stronger feedbacks onto that mode. Since these same feedbacks are expected to47

be involved in the jet response to forcings, it makes sense that a jet with greater SAM per-48

sistence shifts further under forcing. Such reasoning follows the theoretical arguments of49

the Fluctuation-Dissipation theorem (FDT) [Leith, 1975] and it is assumed that the forced50

response is well predicted by the character of the dominant mode of natural variability51

(in this case the SAM).52

Here, building on the work of KG2010, we analyze a large set of models from the Coupled53

Model Intercomparison Project, Phase 5 (CMIP5), and revisit the relationships between54

jet latitude, forced response, and the SAM timescale. Our focus is on seasonality and55

we find that the correlation between jet latitude and forced response is strong in winter56

(JJA) but insignificant during the summer months (DJF). In contrast, the model spread in57

SAM timescales displays the opposite seasonality, being minimal in JJA. The previously58

reported relationships between annual mean jet response and annual mean SAM timescale59

in global climate models are, therefore, mixing relationships in different seasons. This60

questions the appropriateness of interpretations following the FDT, highlights the present61

lack of understanding of the fundamental processes involved and motivates future work62

that should aim to remedy this.63

2. Data and Diagnostics

2.1. Data

We use data from 35 models that participated in CMIP5 (supplementary Table 1). Our64

focus is primarily on the historical and RCP8.5 runs, defining the “Past” as 1979-2005 of65

D R A F T February 25, 2016, 10:00am D R A F T

Page 6: Revisiting the relationship between jet position, · 1 Revisiting the relationship between jet position, 2 forced response and annular mode variability in the 3 southern mid-latitudes

X - 6 SIMPSON & POLVANI: SH JET LATITUDES, JET SHIFTS AND THE SAM

historical, and the “Future” as 2070-2099 of RCP8.5. A complementary analysis is also66

performed using a 200 year sample from the piControl runs, and years 50 to 100 of the67

abrupt4xCO2 runs, to eliminate uncertainties that may be introduced by varying model68

representations of forcings other than CO2. The primary field of interest is the 700hPa69

zonal mean zonal wind u, but we also use 500hPa zonal mean geopotential height Φ, for the70

calculation of the SAM timescales (for consistency with the majority of previous studies71

on SAM persistence [e.g., Baldwin et al. , 2003; Gerber et al., 2010; Kidston and Gerber ,72

2010]). These fields are first interpolated onto a common 2o latitude grid. ERA-Interim73

reanalysis data [Dee et al., 2011] from 1979 to 2005 are also used.74

2.2. Diagnostics

The jet “response” is characterized by both simple geometric parameters, and its rela-75

tion to the first two Empirical Orthogonal Functions (EOFs) of natural variability. The76

geometric parameters are depicted in Fig. 1 (a). Jet latitude is defined as the location of77

the maximum 700hPa u in the SH, obtained by a quadratic fit using the maximum grid78

point and two points either side. The Past and Future jet latitudes are denoted φo and79

φo +∆φ respectively, so that ∆φ is the future jet shift. The Future-Past difference in u80

typically consists of a dipolar structure, producing a poleward shift (Fig. 1a, green line).81

The peak and trough latitudes of this structure are denoted φP and φT , and the peak and82

trough magnitudes are denoted P and T .83

The structure of the natural variability of the jet is characterized by the first two84

EOFs of Past daily u, for the 23 models with daily data available (supplementary Table85

1). These EOFs are calculated for the DJF and JJA seasons separately, using 20oS to86

D R A F T February 25, 2016, 10:00am D R A F T

Page 7: Revisiting the relationship between jet position, · 1 Revisiting the relationship between jet position, 2 forced response and annular mode variability in the 3 southern mid-latitudes

SIMPSON & POLVANI: SH JET LATITUDES, JET SHIFTS AND THE SAM X - 7

70oS de-trended and de-seasonalized u, following Baldwin et al. [2009]. The EOF1 and87

EOF2 wind anomalies as a function of latitude, φ, for each model, i, are denoted u1(φ, i)88

and u2(φ, i) and these are normalized such that the maximum anomaly is 1ms−1. An89

illustrative example is shown in Fig. 1b: EOF1 (i.e., the SAM) typically represents90

latitudinal shifts of the jet, whereas EOF2 typically typically represents variations in jet91

speed. The peak and trough latitudes of EOF1 are denoted φ1+ and φ1−, and the peak92

latitude of EOF2 is denoted φ2.93

The SAM timescale (τ) calculation is performed using 500hPa Φ following Gerber et al.94

[2010]. In brief, the autocorrelation function (ACF) of the first EOF of daily Φ is calcu-95

lated after removing the global mean, de-seasonalizing and de-trending. Unlike for the96

u EOFs above, the EOF here is defined using all days of the year, for consistency with97

previous studies. The timescale for each day of the year is the 1/e-folding timescale of98

this ACF after smoothing over a 181 day window with a Gaussian filter with full width99

half maximum of 42 days. Seasonal or annual averages of these smoothed daily values of100

τ are then computed.101

3. Results

3.1. The relationship between ∆φ, φo and τ

We start by extending the findings of KG2010 to CMIP5, and show in Figs. 2a-c, the102

annual mean relationships of ∆φ and τ with φo (denoted Cor(φo,∆φ) and Cor(φo, τ)), as103

well as the correlation between ∆φ and τ . As already noted by Wilcox et al. [2012] with a104

smaller subset, the CMIP5 models exhibit similar behavior to those examined by KG2010:105

lower latitude jets shift further poleward in the future (Fig. 2a) and exhibit greater SAM106

D R A F T February 25, 2016, 10:00am D R A F T

Page 8: Revisiting the relationship between jet position, · 1 Revisiting the relationship between jet position, 2 forced response and annular mode variability in the 3 southern mid-latitudes

X - 8 SIMPSON & POLVANI: SH JET LATITUDES, JET SHIFTS AND THE SAM

timescales (Fig. 2b). Note that these correlations are significant. The correlation between107

∆φ and τ is not statistically significant (Fig. 2c), although it is of the same sign reported108

by KG2010, i.e, models with greater SAM timescales exhibit larger jet shifts.109

These relationships are then shown separately for DJF and JJA in Figs. 2d-i.110

Cor(φo,∆φ) is much stronger in JJA (panel g) than in DJF (panel d). In fact, it smoothly111

evolves between a minimum in summer (DJF) and a maximum in winter (JJA) (Fig. 3).112

This seasonality was also noted by KG2010, who speculated that the reduced correlation113

in DJF may be due to disparity in the representation of stratospheric ozone forcing among114

models [Son et al., 2008]. However, it persists in CMIP5, despite the more standardized115

treatment of ozone depletion [Eyring et al., 2013]. To further confirm that this season-116

ality occurs regardless of the representation of stratospheric ozone, the green points in117

Fig. 3 show that for the abrupt4xCO2− piControl runs, Cor(φo,∆φ) displays the same118

seasonality.119

If the aforementioned FDT arguments explain Cor(φo,∆φ), then a similar seasonal120

dependence of the relationship between τ and φo would be expected. However, quite the121

opposite is seen. While Cor(φo, τ) is significant in both DJF and JJA, it is clear that122

the majority of the variance in τ , present in the annual mean, comes from DJF (Fig. 3,123

and compare Figs. 2e and h). During JJA, τ is fairly constant across models, and much124

closer to the reanalysis, with the slope of the best fitting regression line between τ and φo125

being an order of magnitude less than in DJF. Versions of Figs. 2 (h) and (i) with more126

appropriate axes for viewing the model spread in JJA are shown in supplementary Figure127

1, making clear that models at opposite ends of the φo, and hence ∆φ, range, exhibit128

D R A F T February 25, 2016, 10:00am D R A F T

Page 9: Revisiting the relationship between jet position, · 1 Revisiting the relationship between jet position, 2 forced response and annular mode variability in the 3 southern mid-latitudes

SIMPSON & POLVANI: SH JET LATITUDES, JET SHIFTS AND THE SAM X - 9

virtually identical values of τ . Hence, the annual relationships in panels a-c are mixing129

strong relationships between ∆φ and φo in JJA and τ and φo in DJF.130

It is worth cautioning that the SAM timescale is not solely indicative of the strength131

of feedback processes, as it can be influenced by external intraseasonal drivers of jet132

variability [Keeley et al., 2009; Simpson et al., 2011]. Ideally, one would want to use a133

directly calculated measure of feedback strength, such as proposed in Simpson et al. [2013],134

rather than τ , but the short timescales in JJA make the feedback strength calculation135

very uncertain in those months, so we make do with τ as a proxy. External drivers of136

variability certainly inflate the timescales in DJF compared to JJA [Keeley et al., 2009;137

Simpson et al., 2011], which could account for some of the increased variance in τ then,138

but that is unlikely to be the whole story. DJF is the primary season where the eddy139

feedback strength exhibits considerable spread across the models and is biased relative to140

the reanalysis [Simpson et al., 2013, their Fig 10].141

The seasonal decomposition, therefore, indicates that our understanding of the SH jet142

response to forcing, and its connection to jet variability and climatological jet position,143

is fundamentally incomplete. An improved understanding of the relevant processes is144

needed, if we are to explain this seasonality. In the season when lower latitude jets145

respond more to forcings (JJA), all models exhibit similar SAM timescales (and thus146

inferred eddy feedbacks), making it difficult to argue that lower latitude jets shift more147

because of stronger eddy feedbacks. It is not clear why Cor(φo,∆φ) should be so strong148

in JJA and we do not provide a complete answer below, but we do shed light on other149

seasonal dependencies that are likely part of the story.150

D R A F T February 25, 2016, 10:00am D R A F T

Page 10: Revisiting the relationship between jet position, · 1 Revisiting the relationship between jet position, 2 forced response and annular mode variability in the 3 southern mid-latitudes

X - 10 SIMPSON & POLVANI: SH JET LATITUDES, JET SHIFTS AND THE SAM

3.2. The latitudinal structure and magnitude of the jet response to forcing

Considering the properties of the zonal wind response to forcing depicted schematically151

in Fig. 1a, two components could contribute to the jet latitude/jet shift relationship152

during JJA: a “structural” component and a “magnitude” component. The structural153

component refers to the fact that the latitude of the wind responses (φT and φP ) relative154

to the climatological jet may depend on jet latitude in such a way as to represent more155

of a poleward shift for lower latitude jets. The magnitude component refers to the fact156

that the magnitude of the wind response (P and T ) may be greater for the lower latitude157

jets. The potential for contributions from these two components to the jet latitude/jet158

shift relationship is explored through Figs. 4a-d.159

Consider first the structural component, i.e. how φT and φP vary with φo. Contrasting160

Figs. 4a and b, there is a clear difference between DJF and JJA. In DJF, φP and φT are161

centered roughly around φo i.e., the wind response represents a poleward shift, regardless162

of climatological jet position. This is not so in JJA. Rather than being equidistant from163

φo, φT is roughly 30oS and φP is roughly 50oS regardless of where the climatological164

jet is located. As a result, the response transitions from a shift to a strengthening as165

one considers higher latitude jets in JJA. This was previously reported by Barnes and166

Polvani [2013] for the annual mean, but appears to be a strongly seasonal effect, occurring167

primarily in the winter.168

Next, consider the magnitude component of the response, depicted in Figs. 4c and d,169

which show P and T versus φo. In DJF, neither is correlated with φo, whereas in JJA, the170

correlations are significant: lower latitude jets exhibit larger wind responses. So, both the171

D R A F T February 25, 2016, 10:00am D R A F T

Page 11: Revisiting the relationship between jet position, · 1 Revisiting the relationship between jet position, 2 forced response and annular mode variability in the 3 southern mid-latitudes

SIMPSON & POLVANI: SH JET LATITUDES, JET SHIFTS AND THE SAM X - 11

structural and the magnitude component may be contributing to the strong relationship172

between ∆φ and φo in JJA, whereas neither is present in DJF.173

3.3. The relation of the response to the dominant modes of variability

The response to a forcing often bears a strong resemblance to the dominant modes174

of natural variability. Indeed, this is one of the key assumptions underlying the FDT175

arguments previously proposed as an explanation of the jet shift/jet latitude correlation.176

It is, therefore, instructive to examine how the forced response, in each season, relates to177

the dominant modes of variability. Furthermore, as will be shown below, such an analysis178

allows for quantification of the relative importance of the “magnitude” and “structural”179

components, discussed above.180

In Figs. 4e and f we illustrate how the structure of the first two EOFs of natural vari-181

ability depend on the jet latitude, through the parameters φ1+, φ1− and φ2 (see Fig. 1b).182

In DJF, much like the forced response in the panels above, EOF1 always represents a183

latitudinal shift, as the red/blue pairs follow φo. Similarly, EOF2 (in green) is always a184

strengthening, since φ2 falls very much on the 1:1 line. In contrast, and similar again to185

the forced response in the panels above, in JJA the EOF parameters change with φ0 such186

that EOF1 and EOF2 cannot be said to always represent a clear shift and strengthening,187

respectively (the colored points have a shallower slope than the black 1:1 line). This too188

was previously reported by Barnes and Polvani [2013] for the annual mean but, again, is189

primarily a JJA effect.190

Next we relate the forced response of the jet to these modes of variability. First,191

we synthetically construct the Future winds uF (φ, i) using the Past winds (uP ) and the192

D R A F T February 25, 2016, 10:00am D R A F T

Page 12: Revisiting the relationship between jet position, · 1 Revisiting the relationship between jet position, 2 forced response and annular mode variability in the 3 southern mid-latitudes

X - 12 SIMPSON & POLVANI: SH JET LATITUDES, JET SHIFTS AND THE SAM

projection of the wind response onto these two EOFS via the expression:193

uF (φ, i) = uP (φ, i) +K1(i)u1(φ, i) +K2(i)u2(φ, i), (1)194

where, for each model i, uF is the constructed Future wind and K1(i) and K2(i) are195

the projections of the response onto the normalized first and second EOFs (u1 and u2).196

Figs. 5a and e demonstrate that this reconstruction works exceedingly well for capturing197

∆φ (except for 2 models in JJA, shown by open circles, that are omitted from the following198

analysis).199

Next, uF is decomposed into contributions from the multi-model mean projection onto200

the EOFs (K1 and K2) and the deviations therefrom (K ′

1 and K ′

2) i.e.,201

uF (φ, i) = uP (φ, i) +K1u1(φ, i) +K2u2(φ, i) +K ′

1(i)u1(φ, i) +K ′

2(i)u2(φ, i) (2)202

In this framework, the FDT arguments would hold if models which longer SAM timescales203

also possessed larger values ofK ′

1. In DJF, the wind response almost entirely projects onto204

EOF 1 (Fig. 5b) but, while lower latitude jets exhibit longer SAM timescales (Fig. 2e), K ′

1205

is completely uncorrelated with φo (Fig. 5c). The behavior in JJA is rather different: the206

wind response projects onto both EOFs 1 and 2 (Fig. 5f). Furthermore, the magnitude of207

the projection onto EOF 1 is significantly correlated with φo (Fig. 5g); which is another208

way of capturing the magnitude component of the response already illustrated in Fig. 4d.209

One advantage of using the synthetically constructed response is that it allows for210

quantification of the importance of the magnitude and structural components. The former211

arises from the 4th and 5th terms in Eq. 2, so omitting those terms in the construction,212

brings out the relationship between φo and ∆φ that arises from the purely structural213

component alone i.e., the component due to the differing position of the wind anomalies214

D R A F T February 25, 2016, 10:00am D R A F T

Page 13: Revisiting the relationship between jet position, · 1 Revisiting the relationship between jet position, 2 forced response and annular mode variability in the 3 southern mid-latitudes

SIMPSON & POLVANI: SH JET LATITUDES, JET SHIFTS AND THE SAM X - 13

relative to φo, here brought about by the differing structures of u1 and u2. This gives is215

an idea of the influence of the structural effect, but the complete jet shift is not a simple216

sum of this and that due to the 4th and 5th terms, due to non-linearities. Figs. 5d and h217

show that in DJF the structural component does not contribute to the dependency of ∆φ218

on φo, whereas in JJA it explains a good fraction of the dependency. For all models, in219

JJA, the regression of ∆φ on φo yields a slope of -0.35◦/◦ (Fig. 2g) and the smaller model220

subset in Fig. 5 yields a similar slope (-0.34◦/◦). When the structural component alone is221

considered the slope is -0.18◦/◦ (Fig.5h) allowing us to conclude that the magnitude and222

structural components are of roughly comparable importance in contributing to the jet223

latitude/jet shift relationship in this season. This highlights the fact that the situation is224

rather complex, as the wind response in JJA does not project uniquely onto EOF1, nor225

does it vary across models according to the magnitude component alone, as one might226

have hoped if the FDT arguments were to apply in a simple way.227

4. Discussion and Conclusions

By revisiting the relationship between each model’s climatological jet latitude, forced228

response and SAM timescale in the CMIP5 dataset, we have demonstrated that a strong229

seasonality underlies previously reported relationships. In the annual mean, lower latitude230

jets (a) shift further poleward when forced and (b) exhibit greater SAM timescales in231

their current climatology. Arguments such as those of the FDT would apply if these two232

relationships occurred in the same seasons, but they don’t. The CMIP5 models reveal233

that:234

D R A F T February 25, 2016, 10:00am D R A F T

Page 14: Revisiting the relationship between jet position, · 1 Revisiting the relationship between jet position, 2 forced response and annular mode variability in the 3 southern mid-latitudes

X - 14 SIMPSON & POLVANI: SH JET LATITUDES, JET SHIFTS AND THE SAM

• The correlation between jet latitude and jet shift is strong, only during the extended235

winter season (see Fig 3). The usefulness of this relationship as an emergent constraint236

[Collins et al., 2012] on future projections is, therefore, restricted to this season.237

• While the correlation between jet latitude and jet shift is very strong during JJA,238

models exhibit very similar SAM timescales in that season, making it difficult to argue that239

the jet latitude/jet shift relationship is brought about through a latitudinal dependence240

of eddy/mean flow feedback strength, inferred from SAM timescales.241

Two components, of comparable importance, have been shown to give rise to the strong242

relationship between φo and ∆φ in JJA: (1) the positioning of the wind response relative243

to the climatological jet varies with jet location such that it represents less of a poleward244

shift for higher latitude jets (Fig. 4b) and (2) the magnitude of the wind response is245

larger for lower latitude jets (Fig. 4d). However, the reason behind these two components246

is not known at present. Mechanisms have been proposed, through simplified modeling,247

to explain a relationship between jet latitude and jet shift [Barnes et al., 2010; Simpson248

et al., 2012] and whether these are relevant to the JJA relationship in comprehensive249

models is unclear, since accompanying a greater jet shift is a more persistent annular250

mode in these idealized studies, an aspect that is not seen in the more complex models251

analyzed here.252

In summary, much remains to be understood about the relationships between a model’s253

climatological jet position, response to forcing and annular mode variability. While this254

study perhaps raises more questions than provides answers, it points toward future re-255

search directions that should be taken. An improved understanding of the seasonalities256

D R A F T February 25, 2016, 10:00am D R A F T

Page 15: Revisiting the relationship between jet position, · 1 Revisiting the relationship between jet position, 2 forced response and annular mode variability in the 3 southern mid-latitudes

SIMPSON & POLVANI: SH JET LATITUDES, JET SHIFTS AND THE SAM X - 15

identified here should lead to both interesting insights into the dynamics of the SH jet257

stream as well as an improved understanding of model limitations and their impact on258

predictions of the SH circulation.259

Acknowledgments. The National Center for Atmospheric Research is sponsored by260

the National Science Foundation (NSF). This work was also partially supported by NSF261

award AGS-1317469 and the work of LMP is funded, in part, by an NSF grant to Columbia262

University. We acknowledge the World Climate Research Programme’s Working Group263

on Coupled Modelling, which is responsible for CMIP, and we thank the climate mod-264

eling groups (listed in Table 1 of this paper) for producing and making available their265

model output. For CMIP the U.S. Department of Energy’s Program for Climate Model266

Diagnosis and Intercomparison provides coordinating support and led development of soft-267

ware infrastructure in partnership with the Global Organization for Earth System Science268

Portals.269

References

Baldwin, M. P., D. B. Stephenson, D. W. J. Thompson, T. J. . Dumkerton, A. J. Charlton270

and A. J. O’Neil (2003), Stratospheric Memory and Skill of Extended-Range Weather271

Forecasts, Science, 301, 636–640.272

Baldwin, M. P., D. B. Stephenson, and I. T. Jollife (2009), Spatial weighting and iterative273

projection methods for eofs, J. Clim., 22, 234–243.274

Barnes, E. A., and L. Polvani (2013), Response of the midlatitude jets, and of their275

variability, to increased greenhouse gases in the cmip5 models, J. Clim., 26, 7117–7135.276

D R A F T February 25, 2016, 10:00am D R A F T

Page 16: Revisiting the relationship between jet position, · 1 Revisiting the relationship between jet position, 2 forced response and annular mode variability in the 3 southern mid-latitudes

X - 16 SIMPSON & POLVANI: SH JET LATITUDES, JET SHIFTS AND THE SAM

Barnes, E. A., D. L. Hartmann, D. M. W. Frierson, and J. Kidston (2010), Effect of277

latitude on the persistence of eddy-driven jets, Geophys. Res. Lett., 37, L11,804278

Biastoch, A., C. W. Boning, F. U. Schwarzkopf, and J. R. E. Lutjeharms (2009), Increase279

in agulhas leakage due to poleward shift of southern hemisphere westerlies, Nature, 462,280

495–499.281

Bracegirdle, T. J., E. Shuckburgh, J.-B. Sallee, Z. Wang, A. J. S. Meijers, N. Bruneau,282

T. Phillips, and L. J. Wilcox (2013), Assessment of surface winds over the atlantic,283

indian and pacific ocean sectors of the southern ocean in cmip5 models: historical bias,284

forcing response and state dependence., J. Geophys. Res. Atm., 118285

Collins, M., R. E. Chandler, P. M. Cox, J. M. Huthnance, J. Rougier, and D. B. Stephen-286

son (2012), Quantifying future climate change., Nat. Clim. Change., 2287

Dee, D. P., S. M. Uppala, a. J. Simmons, P. Berrisford, P. Poli, S. Kobayashi, U. Andrae,288

M. a. Balmaseda, G. Balsamo, P. Bauer, P. Bechtold, a. C. M. Beljaars, L. van de Berg,289

J. Bidlot, N. Bormann, C. Delsol, R. Dragani, M. Fuentes, a. J. Geer, L. Haimberger,290

S. B. Healy, H. Hersbach, E. V. Holm, L. Isaksen, P. Ka llberg, M. Kohler, M. Ma-291

tricardi, a. P. McNally, B. M. Monge-Sanz, J.-J. Morcrette, B.-K. Park, C. Peubey,292

P. de Rosnay, C. Tavolato, J.-N. Thepaut, and F. Vitart (2011), The ERA-Interim re-293

analysis: configuration and performance of the data assimilation system, QJRMS, 137,294

553–597295

Devore, J. L. (1999), Probability and Statistics for Engineering and the Sciences, 5th Ed,296

535-536 pp., Brooks/Cole.297

D R A F T February 25, 2016, 10:00am D R A F T

Page 17: Revisiting the relationship between jet position, · 1 Revisiting the relationship between jet position, 2 forced response and annular mode variability in the 3 southern mid-latitudes

SIMPSON & POLVANI: SH JET LATITUDES, JET SHIFTS AND THE SAM X - 17

Eyring, V., J. Arblaster, I. Cionni, J. Sedlacek, J. Perlwitz, P. Young, S. Bekki,298

D. Bergmann, P. Cameron-Smith, W. J. Collins, et al. (2013), Long-term ozone changes299

and associated climate impacts in cmip5 simulations, Journal of Geophysical Research:300

Atmospheres, 118 (10), 5029–5060.301

Fyfe, J. C., and O. A. Saenko (2006), Simulated changes in the extratropical Southern302

Hemisphere winds and currents, Geophys. Res. Lett., 33, L06,701303

Gent, P. (2016), Effects of southern hemisphere wind changes on the meridional overturn-304

ing circulation in models, Ann. Rev. Mar. Sci., 8, 2.1–2.16305

Gerber, E. P., and G. K. Vallis (2007), Eddy-Zonal Flow Interactions and the Persistence306

of the Zonal Index, J. Atmos. Sci., 64, 3296–3311.307

Gerber, E. P., M. P. Baldwin, H. Akiyoshi, J. Austin, S. Bekki, P. Braesicke, N. Butchart,308

M. Chipperfield, M. Dameris, S. Dhomse, S. M. Frith, R. R. Garcia, H. Garny, A. Get-309

telman, S. C. Hardiman, A. Karpechko, M. Marchand, O. Morgenstern, J. E. Nielsen,310

S. Pawson, T. Peter, D. a. Plummer, J. a. Pyle, E. Rozanov, J. F. Scinocca, T. G.311

Shepherd, and D. Smale (2010), Stratosphere-troposphere coupling and annular mode312

variability in chemistry-climate models, J. Geophys. Res., 115, D00MN06313

Keeley, S. P. E., R. T. Sutton, and L. C. Shaffrey (2009), Does the north atlantic oscillation314

show unusual persistence on intraseasonal timescales?, Geophys. Res. Lett., 36, L22706315

Kidston, J., and E. P. Gerber (2010), Intermodel variability of the poleward shift of the316

austral jet stream in the CMIP3 integrations linked to biases in 20th century climatol-317

ogy, Geophys. Res. Lett., 37, L09,708318

D R A F T February 25, 2016, 10:00am D R A F T

Page 18: Revisiting the relationship between jet position, · 1 Revisiting the relationship between jet position, 2 forced response and annular mode variability in the 3 southern mid-latitudes

X - 18 SIMPSON & POLVANI: SH JET LATITUDES, JET SHIFTS AND THE SAM

Leith, C. E. (1975), Climate response and fluctuation dissipation, J. Atmos. Sci., 32,319

2022–2026.320

Marshall, J., and K. Speer (2012), Closure of the meridional overturning circulation321

through southern ocean upwelling, Nature Geoscience, 5 (3), 171–180.322

Sigmond, M., and J. C. Fyfe (2010), Has the ozone hole contributed to increased Antarctic323

sea ice extent?, Geophys. Res. Lett., 37, L18,502324

Simpson, I. R., M. Blackburn, J. D. Haigh, and S. N. Sparrow (2010), The impact of the325

state of the troposphere on the response to stratospheric heating in a simplified GCM.,326

J. Clim., 23, 6166–6185.327

Simpson, I. R., P. Hitchcock, T. G. Shepherd, and J. F. Scinocca (2011), Stratospheric328

variability and tropospheric annular mode timescales, Geophys. Res. Lett., 38, L20806.329

Simpson, I. R., M. Blackburn, and J. D. Haigh (2012), A mechanism for the effect of330

tropospheric jet structure on the annular mode-like repsonse to stratospheric forcing,331

J. Atmos. Sci., 69, 2152–2170.332

Simpson, I. R., T. G. Shepherd, P. Hitchcock, and J. F. Scinocca (2013), Southern annular333

mode dynamics in observations and models, part 2: Eddy feedbacks, J. Clim., 26, 5220–334

5241.335

Smith, K. L., L. M. Polvani, and D. R. Marsh (2012), Mitigation of 21st century antarctic336

sea ice loss by stratospheric ozone recovery, Geophys. Res. Lett., 39, L20701.337

Son, S., E. P. Gerber, J. Perlwitz, L. M. Polvani, and N. Gillett (2010), The Impact of338

Stratospheric Ozone on the Southern Hemisphere Circulation Changes : A Multimodel339

Assessment, J. Atmos. Sci., 115, 1–55.340

D R A F T February 25, 2016, 10:00am D R A F T

Page 19: Revisiting the relationship between jet position, · 1 Revisiting the relationship between jet position, 2 forced response and annular mode variability in the 3 southern mid-latitudes

SIMPSON & POLVANI: SH JET LATITUDES, JET SHIFTS AND THE SAM X - 19

Son, S.-W., and S. Lee (2005), The Response of Westerly Jets to Thermal Driving in a341

Primitive Equation Model, J. Atmos. Sci., 62, 3741–3757.342

Son, S.-W., L. Polvani, D. Waugh, H. Akiyoshi, R. Garcia, D. Kinnison, S. Pawson,343

E. Rozanov, T. Shepherd, and K. Shibata (2008), The impact of stratospheric ozone344

recovery on the southern hemisphere westerly jet, Science, 320 (5882), 1486–1489.345

Swart, N. C., and J. C. Fyfe (2012), Observed and simulated changes in southern hemi-346

sphere surface westerlies, Geophys. Res. Lett., 39, L16,711347

Wilcox, L. H., A. J. Charlton-Perez, and L. J. Gray (2012), Trends in austral jet position348

in ensembles of high- and low-top cmip5 models, J. Geophys. Res., 117, D13115.349

(a) Jet response

70S 60S 50S 40S 30S 20SLatitude

0

5

10

15

20

700h

Pa

zona

l mea

n u

(m/s

)

φP

φT

φo

∆φ

P

T

PastFutureFuture-Pastresponse

(b) Jet variability

70S 60S 50S 40S 30SLatitude

-1.0

-0.5

0.0

0.5

1.0

1.5

2.0

2.5

700h

Pa

zona

l mea

n u

(m/s

)

φEOF1+ φEOF1-φEOF2

Past/8EOF1

EOF2

Figure 1. An illustrative example (DJF, CanESM2) of parameters used. (a) Geometric param-

eters characterizing Past and Future climatologies, (b) parameters characterizing the structure

of the Past EOF1 and EOF2. In (a) φo = past jet latitude, ∆φ=Future-Past jet shift, φP and

φT = locations of the peak and trough of the Future-Past difference dipole and P and T are their

magnitudes. In (b) φEOF1+ and φEOF1− are the peak and trough latitudes of EOF1 and φEOF2

the peak latitude of EOF2 and the structure of the past climatological jet is shown for reference.

D R A F T February 25, 2016, 10:00am D R A F T

Page 20: Revisiting the relationship between jet position, · 1 Revisiting the relationship between jet position, 2 forced response and annular mode variability in the 3 southern mid-latitudes

X - 20 SIMPSON & POLVANI: SH JET LATITUDES, JET SHIFTS AND THE SAM

(a) Annual

-6

-4

-2

0

2

∆φ

r=-0.55 (-0.75,-0.27)m=-0.21

∆φ vs ττ vs φo∆φ vs φo

(b) Annual

0

10

20

30

40

50

6070

τ (d

ays)

r=0.60 (0.24,0.81)m=+1.57

(c) Annual

-6

-4

-2

0

2

∆φ

r=-0.38 (-0.68,0.04)m=-0.06

ACCESS1-0ACCESS1-3bcc-csm1-1bcc-csm1-1-mBNU-ESMCanESM2CCSM4

CESM1-CAM5CESM1-WACCMCMCC-CMCMCC-CMSCNRM-CM5CSIRO-Mk3-6-0FGOALS-g2

FIO-ESMGFDL-CM3GFDL-ESM2GGFDL-ESM2MGISS-E2-HGISS-E2-RHadGEM2-AO

HadGEM2-CCHadGEM2-ESinmcm4IPSL-CM5A-LRIPSL-CM5A-MRIPSL-CM5B-LRMIROC5

MIROC-ESMMIROC-ESM-CHEMMPI-ESM-LRMPI-ESM-MRMRI-CGCM3NorESM1-MNorESM1-ME

(d) DJF

-6

-4

-2

0

2

∆φ

r=-0.22 (-0.52,0.12)m=-0.11

(e) DJF

0

10

20

30

40

50

6070

τ (d

ays)

r=0.59 (0.24,0.81)m=+3.40

(f) DJF

-6

-4

-2

0

2

∆φ

r=-0.14 (-0.52,0.29)m=-0.01

(g) JJA

55S 50S 45S 40S 35Sφo

-6

-4

-2

0

2

∆φ

r=-0.80 (-0.90,-0.64)m=-0.35

(h) JJA

55S 50S 45S 40S 35Sφo

010

20

30

40

50

6070

τ (d

ays)

r=0.54 (0.17,0.78)m=+0.34

(i) JJA

0 10 20 30 40 50 60 70τ (days)

-6

-4

-2

0

2

∆φ

r=-0.39 (-0.69,0.03)m=-0.26

Figure 2. (left) ∆φ versus φo, (middle) τ versus φo and (right) ∆φ versus τ . (a)-(c) annual

means, (d)-(f) DJF and (g)-(i) JJA. Correlation coefficients (r) and their 95% confidence interval

calculated using the Fisher transform [Devore, 1999] are quoted along with the best fitting regres-

sion slope (m) (in grey where the confidence interval on the correlation coefficient encompasses

zero). Dashed lines depict ERA-Interim values where appropriate.

D R A F T February 25, 2016, 10:00am D R A F T

Page 21: Revisiting the relationship between jet position, · 1 Revisiting the relationship between jet position, 2 forced response and annular mode variability in the 3 southern mid-latitudes

SIMPSON & POLVANI: SH JET LATITUDES, JET SHIFTS AND THE SAM X - 21

Seasonal variation in Cor(φo, ∆φ) and the standard deviation of SAM timescales

jan feb mar apr may jun jul aug sep oct nov dec

RCP8.5 - historical 4xCO2 - 1850 σ(τ)-0.2

0

0.2

0.4

0.6

0.8

1.0

Cor

(φo,

∆φ)

0

5

10

15

σ(τ)

Figure 3. Seasonal cycle of monthly Cor(φo,∆φ) (black and green) and the standard devia-

tion, across models, of monthly averaged τ (red). Black points: Future=RCP8.5, Past=historical.

Green points: Future=abrupt4xCO2, Past=piControl. Error bars display 95% confidence inter-

vals on Cor(φo,∆φ) calculated using the Fisher transform [Devore, 1999].

D R A F T February 25, 2016, 10:00am D R A F T

Page 22: Revisiting the relationship between jet position, · 1 Revisiting the relationship between jet position, 2 forced response and annular mode variability in the 3 southern mid-latitudes

X - 22 SIMPSON & POLVANI: SH JET LATITUDES, JET SHIFTS AND THE SAM

(a) φP and φT vs φo

55S 50S 45S 40S 35Sφo

60S

50S

40S

30S

20S

10S

φ P a

nd φ

T

φP

φTr=0.80 (0.63,0.90)

r=0.84 (0.70,0.92)

DJF JJA(b) φP and φT vs φo

55S 50S 45S 40S 35Sφo

60S

50S

40S

30S

20S

10S

φ P a

nd φ

T

φP

φTr=0.15 (-0.20,0.46)

r=0.24 (-0.10,0.54)

(c) P and T vs φo

55S 50S 45S 40S 35Sφo

-4

-2

0

2

4

6

P a

nd T

PT

r=0.27 (-0.16,0.51)

r=-0.19 (-0.49,0.19)

(d) P and T vs φo

55S 50S 45S 40S 35Sφo

-4

-2

0

2

4

6

P a

nd T

PT

r=0.64 (0.35,0.79)

r=-0.56 (-0.74,-0.24)

(e) φEOF1+, φEOF1- and φEOF2 vs φo

55S 50S 45S 40S 35Sφo

60S

50S

40S

30S

20S

10S φEOF1+φEOF1-φEOF2

r=0.94 (0.87,0.98)r=0.95 (0.89,0.98)r=0.94 (0.86,0.97)

(f) φEOF1+, φEOF1- and φEOF2 vs φo

55S 50S 45S 40S 35Sφo

60S

50S

40S

30S

20S

10Sr=0.77 (0.52,0.90)r=0.69 (0.39,0.86)r=0.74 (0.47,0.88)

φEOF1+φEOF1-φEOF2

Figure 4. (left) DJF and (right) JJA. (a)-(d) relationships between geometric parameters

of the RCP8.5-historical difference and φo. ((a) and (b)) φP and φT vs φo, ((c) and (d)) P

and T vs φo. Circled points are considered outliers due to weak responses, leading to difficulty

in characterizing the response by the representative dipole structure. These are omitted in the

calculation of correlation coefficients. ((e) and (f)) φEOF1−, φEOF1+ and φEOF2 vs φo. Correlation

coefficients and their 95% confidence interval are quoted (in grey where statistically insignificant).

Solid lines in (a),(b),(e) and (f) depict the 1:1 line.D R A F T February 25, 2016, 10:00am D R A F T

Page 23: Revisiting the relationship between jet position, · 1 Revisiting the relationship between jet position, 2 forced response and annular mode variability in the 3 southern mid-latitudes

SIMPSON & POLVANI: SH JET LATITUDES, JET SHIFTS AND THE SAM X - 23

(a) constructed ∆φ vs actual ∆φ

-12 -10 -8 -6 -4 -2 0 2Actual ∆φ

-12

-10

-8

-6

-4

-2

0

2

Con

stru

cted

∆φ

(b) K1 and K2

-0.5

0.0

0.5

1.0

1.5

2.0

2.5

K1

or K

2

EOF1 EOF2

(c) Cor(φo,K′)

0.0

0.2

0.4

0.6

Cor

rela

tion

EOF1 EOF2

(d) Structural component

-55 -50 -45 -40 -35φo

-6

-4

-2

0

2

Con

stru

cted

∆φ

m=+0.01r=0.11 (-0.31,0.50)

(e) constructed ∆φ vs actual ∆φ

-12 -10 -8 -6 -4 -2 0 2Actual ∆φ

-12

-10

-8

-6

-4

-2

0

2

Con

stru

cted

∆φ

(f) K1 and K2

-0.5

0.0

0.5

1.0

1.5

2.0

2.5

K1

or K

2

EOF1 EOF2

(g) Cor(φo,K′)

0.0

0.2

0.4

0.6

Cor

rela

tion

EOF1 EOF2

(h) Structural component

-55 -50 -45 -40 -35φo

-6

-4

-2

0

2

Con

stru

cted

∆φ

m=-0.18r=-0.55(-0.79,-0.16)

DJF

JJA

Figure 5. Relating the jet shift to the first two EOFs of natural variability in (top) DJF and

(bottom) JJA. ((a) and (e)) constructed jet shift based on (1) versus actual jet shift. This does

not work well for the two circled models in (e), so these are omitted from the remaining panels.

((b) and (f)) multi-model mean projection of the Future-Past difference onto EOFs 1 and 2, ((c)

and (d)) the correlation between the projection onto EOFs 1 and 2 and φo. Shaded bars are

statistically significant at the 95% level by a students T test in (b) and (f) and by the Fisher

transform in (c) and (g). (d) and (h) The relationship between ∆φ and φo obtained from the

structural effect alone i.e., omitting the last two terms in (2).

D R A F T February 25, 2016, 10:00am D R A F T