El Niño-Southern Oscillation (ENSO) in a Changing Climate ... · 104 the “Climate Diagnostics...

45
1 El Niño-Southern Oscillation (ENSO) in a Changing Climate 1 2 Chapter 10: ENSO Prediction 3 4 Michelle L. L’Heureux 1 , Aaron F.Z. Levine 2 , Matthew Newman 3 , 5 Catherine Ganter 4 , Jing-Jia Luo 5 , Michael K. Tippett 6 , Timothy N. Stockdale 7 6 7 1 National Oceanic and Atmospheric Administration, NWS/NCEP/Climate Prediction Center, 8 5830 University Research Ct, College Park, MD 20740, USA. 9 10 2 University of Washington, Department of Atmospheric Sciences, Seattle, WA 98195, USA. 11 12 3 University of Colorado/CIRES, NOAA/ESRL Physical Sciences Division, 325 Broadway, 13 Boulder, CO 80305, USA. 14 15 4 Australian Bureau of Meteorology, 700 Collins St, Melbourne, VIC, Australia. 16 17 5 Institute for Climate and Application Research (ICAR)/CICFEM/KLME/ILCEC, Nanjing 18 University of Information Science and Technology, Nanjing, China. 19 20 6 Columbia University, Department of Applied Physics and Applied Mathematics, 500 W. 120th 21 St., New York, NY 10027, USA. 22 23 7 European Centre for Medium-Range Weather Forecasts, Reading, UK. 24 25 Abstract 26 27 The El Niño-Southern Oscillation (ENSO) is a coupled ocean-atmosphere phenomenon 28 of variability that is a leading source of seasonal climate prediction skill across the globe. The 29 first ENSO prediction was made in the mid 1970s, but it was another 10-15 years before 30 operational centers, using simple, coupled climate models, began to make routine ENSO 31 predictions. These early forecast models were succeeded in the 1990s by more sophisticated 32 dynamical and statistical models, which created the basis for real-time seasonal outlooks over the 33 globe. These models, and more recent multi-model ensembles, also inform our understanding 34 and estimates of the predictability and prediction skill of ENSO, which varies seasonally and 35 from decade-to-decade. ENSO predictability largely stems from slowly evolving oceanic 36 conditions, with short-term atmospheric fluctuations often limiting predictability on seasonal 37 timescales. Despite improved models and better initializations, prediction skill remains low for 38 forecasts passing through the boreal spring, the so-called spring prediction barrier. Furthermore, 39

Transcript of El Niño-Southern Oscillation (ENSO) in a Changing Climate ... · 104 the “Climate Diagnostics...

Page 1: El Niño-Southern Oscillation (ENSO) in a Changing Climate ... · 104 the “Climate Diagnostics Bulletin.” Initially, the forecast forum of the Bulletin featured 105 forecasts

1

El Niño-Southern Oscillation (ENSO) in a Changing Climate 1 2

Chapter 10: ENSO Prediction 3 4

Michelle L. L’Heureux1, Aaron F.Z. Levine2, Matthew Newman3, 5 Catherine Ganter4, Jing-Jia Luo5, Michael K. Tippett6, Timothy N. Stockdale7 6

7 1 National Oceanic and Atmospheric Administration, NWS/NCEP/Climate Prediction Center, 8 5830 University Research Ct, College Park, MD 20740, USA. 9 10 2 University of Washington, Department of Atmospheric Sciences, Seattle, WA 98195, USA. 11 12 3 University of Colorado/CIRES, NOAA/ESRL Physical Sciences Division, 325 Broadway, 13 Boulder, CO 80305, USA. 14 15 4 Australian Bureau of Meteorology, 700 Collins St, Melbourne, VIC, Australia. 16 17 5 Institute for Climate and Application Research (ICAR)/CICFEM/KLME/ILCEC, Nanjing 18 University of Information Science and Technology, Nanjing, China. 19 20 6 Columbia University, Department of Applied Physics and Applied Mathematics, 500 W. 120th 21 St., New York, NY 10027, USA. 22 23 7 European Centre for Medium-Range Weather Forecasts, Reading, UK. 24 25 Abstract 26

27

The El Niño-Southern Oscillation (ENSO) is a coupled ocean-atmosphere phenomenon 28

of variability that is a leading source of seasonal climate prediction skill across the globe. The 29

first ENSO prediction was made in the mid 1970s, but it was another 10-15 years before 30

operational centers, using simple, coupled climate models, began to make routine ENSO 31

predictions. These early forecast models were succeeded in the 1990s by more sophisticated 32

dynamical and statistical models, which created the basis for real-time seasonal outlooks over the 33

globe. These models, and more recent multi-model ensembles, also inform our understanding 34

and estimates of the predictability and prediction skill of ENSO, which varies seasonally and 35

from decade-to-decade. ENSO predictability largely stems from slowly evolving oceanic 36

conditions, with short-term atmospheric fluctuations often limiting predictability on seasonal 37

timescales. Despite improved models and better initializations, prediction skill remains low for 38

forecasts passing through the boreal spring, the so-called spring prediction barrier. Furthermore, 39

Page 2: El Niño-Southern Oscillation (ENSO) in a Changing Climate ... · 104 the “Climate Diagnostics Bulletin.” Initially, the forecast forum of the Bulletin featured 105 forecasts

2

prediction skill and predictability have varied significantly over the past couple decades. Higher 40

skill and predictability is evident during periods of larger amplitude (or Eastern Pacific flavored) 41

ENSO events, whereas lower skill/predictability is associated with lower amplitude (or Central 42

Pacific flavored) events. These natural variations in our ability to predict ENSO, together with 43

challenges during 2014-16, motivate the search for understanding of how anthropogenic 44

warming will influence seasonal ENSO prediction. 45

46

13.1. History of ENSO Forecasting 47

48

William Quinn of Oregon State University issued the first known El Niño prediction in 49

1974 (McPhaden et al. 2015; Figure 1). Based on the persistence of the Southern Oscillation 50

Index, he forecasted that a weak El Niño would emerge in early 1975 and cause below-normal 51

precipitation over Indonesia through mid-to-late 1975 (Quinn, 1974a,b). The predicted El Niño 52

did not develop in 1975, and a retrospective analysis offered by McPhaden et al. (2015) indicates 53

that an insufficient build-up of oceanic heat content prevented its development. Regardless, 54

Quinn’s prediction provided a start for ENSO forecasting. Furthermore, this first ENSO 55

prediction illustrates the close relation between seasonal climate prediction and ENSO, which 56

pre-dated a comprehensive understanding of ENSO itself. Sir Gilbert Thomas Walker uncovered 57

the Southern Oscillation in an effort to predict Indian monsoon rainfall (Katz, 2002), and ancient 58

Andean farmers used the visibility of stars in the Pleiades, which is influenced by ENSO, to 59

determine when to plant their crops (Orlove et al., 2000). Even today, the predictability of 60

ENSO offers a scientific basis for seasonal climate outlooks around the globe. 61

Currently, ENSO forecasts are produced by many national meteorological and climate 62

forecast centers around the world. We focus here on the development of operational forecasts, or 63

the regular, timely dissemination of forecasts and data. The operational setting provides a 64

uniquely thorough test of prediction methods, in large part because the future state of ENSO 65

being predicted is truly unknown. The operational scenario is quite different from evaluating 66

model forecasts initialized on historical data when what happened is known. Real-time forecasts 67

test model assumptions on truly independent data, and carry the risk of being exposed as too 68

reliant on past data or based on insufficient or inaccurate theories. Operational time constraints 69

mean that observational data have to be rapidly and accurately collected, assembled, and 70

Page 3: El Niño-Southern Oscillation (ENSO) in a Changing Climate ... · 104 the “Climate Diagnostics Bulletin.” Initially, the forecast forum of the Bulletin featured 105 forecasts

3

ingested to provide the starting conditions for forecast models. Users require regular and routine 71

dissemination of forecasts for decisions and planning at all times, so operational forecasts must 72

be issued even during periods or situations when forecasts might perform poorly. Some of the 73

earliest historical operational developments for ENSO forecasting were made in the United 74

States (National Oceanic and Atmospheric Administration’s Climate Prediction Center, CPC) 75

and Australia (Bureau of Meteorology, BOM). 76

The strong 1982-83 El Niño and its global impacts were well underway by the time 77

scientists noticed it. Its unexpectedly strong development led to accelerated worldwide efforts in 78

climate observations and forecasting. Until this event, ENSO research was not advanced enough 79

to be applied to operational climate outlooks, in part because of the nascent state of climate 80

forecasting. CPC, created just three years earlier (initially as the Climate Analysis Center) to 81

serve as NOAA’s new climate forecast agency, did not issue its first bulletin advertising El Niño 82

until November 1982. This delay was mostly due to inadequate monitoring infrastructure. 83

Satellite monitoring of sea surface temperatures was in place, but contamination by volcanic 84

aerosols from the eruption of El Chichón resulted in sea surface temperatures estimates being too 85

cool. Ship data, on the other hand, showed strong warming in the equatorial Pacific Ocean, but 86

were discounted as erroneous because they were too warm compared to what was expected 87

(Reeves and Gemmill, 2004). These data gaps motivated the establishment of the 10-year 88

Tropical-Ocean-Global-Atmosphere (TOGA) project to improve tropical Pacific Ocean 89

monitoring and ENSO prediction (McPhaden, et al., 2010). In the aftermath of the 1982-83 El 90

Niño, which brought widespread drought to agricultural regions of Australia, Australian farmers 91

were desperate for long-term rainfall forecasts. This in part led to the establishment of the 92

National Climate Centre at BOM and the dedication of resources to improve long-term forecast 93

skill beyond numerical weather timescales. In the wider scientific community, the dramatic 94

nature of the 1982-83 event spurred interest in understanding, modeling and, in time, predicting 95

ENSO variability. 96

At CPC, advances in tropical Pacific monitoring led to the release of the first ENSO 97

Diagnostic Advisory in January 1986 (Figure 2; Reeves and Gemmill, 2004). However, this 98

provided only diagnostic information about current ENSO conditions, which otherwise was still 99

a research and monitoring effort, and forecasts of ENSO were still not routinely issued. Barnett 100

et al. (1988) were among the first to show that forecasts of the 1986-87 El Niño had skill 101

Page 4: El Niño-Southern Oscillation (ENSO) in a Changing Climate ... · 104 the “Climate Diagnostics Bulletin.” Initially, the forecast forum of the Bulletin featured 105 forecasts

4

comparable to operational 30-90 day climate outlooks issued at that time. As a result, in June 102

1989 predictions of ENSO began to be routinely produced at CPC as part of monthly updates of 103

the “Climate Diagnostics Bulletin.” Initially, the forecast forum of the Bulletin featured 104

forecasts of the Niño-3 index from a simplified coupled model of Cane et al. (1986) and a hybrid 105

statistical-dynamical model based on Inoue and O’Brien (1984). With the advent of the TOGA 106

program on prediction (Cane and Sarachik, 1991; Battisti and Sarachick, 1995), the 107

Experimental Long-Lead Forecast Bulletin began in the fall of 1992 to publish forecasts and 108

advances in seasonal prediction from authors throughout the research and operational climate 109

communities. Many other statistical and dynamical models came on board during the 1990s, 110

some of which became the basis for real-time ENSO forecasts and their verifications (Barnston 111

et al., 1994). Ultimately, these advances in modeling driven by ENSO paved the way for 112

operational seasonal outlooks for rainfall and temperature. 113

At BOM, seasonal rainfall outlooks began to be routinely disseminated in 1989 for the 114

“zero-lead”, the period that begins in the month of issuance (for example, a July to September 115

forecast issued in July). These three-category (above, near, or below-average) outlooks for 116

Australian precipitation were based on the ENSO state as measured by the Southern Oscillation 117

Index (SOI). In 1996, the precipitation outlooks were joined by six-monthly ENSO outlooks 118

based on an intermediate coupled model. CPC’s seasonal outlooks were also only for zero-lead 119

until 1995, when increased confidence in ENSO prediction skill led to the extension of United 120

States temperature and precipitation outlooks out to a year (Barnston et al., 1994). However, not 121

until the major 1997-98 El Niño did CPC and BOM produce standalone, operational ENSO 122

discussions issued regularly, even during neutral ENSO conditions. Prior to 1997-98, discussion 123

related to ENSO was embedded within the seasonal climate discussion, and advisories or media 124

releases were only issued when conditions appeared ripe for an event. 125

In Europe, the UK Met Office worked on ENSO modeling with coupled GCMs in the 126

1980s (e.g. Gordon, 1989), but this did not immediately lead to a dynamical ENSO forecasting 127

system: instead, operational seasonal outlooks were first generated by a two-tier system using 128

persisted SST anomalies (Graham et al., 2000). ECMWF started the development of a seasonal 129

forecast system in 1994, and because the 1997-98 El Niño was captured well in their 130

experimental system (Stockdale et al, 1998), they began issuing Niño SST and seasonal forecasts 131

on an operational schedule from December 1997. European groups subsequently collaborated on 132

Page 5: El Niño-Southern Oscillation (ENSO) in a Changing Climate ... · 104 the “Climate Diagnostics Bulletin.” Initially, the forecast forum of the Bulletin featured 105 forecasts

5

the development of multi-model GCM seasonal forecast systems with two research projects 133

(DEMETER, ENSEMBLES), which demonstrated the power of a multi-model approach (Palmer 134

et al, 2004; Weisheimer et al, 2009). This led to the development of the EUROSIP multi-model 135

system, which has produced ENSO and seasonal forecasts operationally since 2005 (Stockdale et 136

al, 2009). A coordinated multi-model GCM approach was also picked up by APCC (the APEC 137

Climate Centre, founded in 2005) in South Korea (Min et al, 2009) and the NMME (North 138

American Multi-Model Ensemble) in North America, starting in 2011 (Kirtman et al, 2014). 139

More recently, EUROSIP has been superseded by the Copernicus Climate Change Service (C3S) 140

multi-model seasonal forecast system. 141

In recent years, there has been an explosion in the number of seasonal climate outlooks 142

produced by coupled dynamical models, all relying upon an expansion of observing networks, 143

improvements in data assimilation, and a tremendous escalation of computing power. While 144

these models make global predictions, ENSO remains a core component of any credible seasonal 145

forecast system. One indicator of its importance is the number of readily available forecasts of 146

ENSO posted online each month. The International Research Institute for Climate and Society 147

(IRI) has the longest-running archive of ENSO model forecasts, but other national and 148

international efforts have resulted in various model displays that are posted in real or near real-149

time. 150

ENSO outlooks often showcase forecasts of SST anomalies in the Niño index regions, 151

spanning the westernmost Niño-4 region to the easternmost Niño-1+2 region. Of these regions, 152

the Niño-3.4 index is often featured because this region has strong correlations with other 153

atmospheric measures of ENSO and with remote climate impacts (Barnston et al., 1997). An 154

expert team was commissioned in the mid-2000s to catalogue El Niño and La Niña definitions 155

used operationally by member countries of the World Meteorological Organization (Horsfall et 156

al., 2006). While a wide variety of indices were used, one of the most common indices was 157

based on 3-month average values in the Niño-3.4 index. Thus, this index is often relied upon for 158

designations of El Niño and La Niña, though the threshold or cut-off can vary by agency. 159

The method or format by which national and international organizations disseminate their 160

ENSO outlooks varies considerably. Some national meteorological services do not routinely 161

write forecast discussions, but instead post forecast model output onto their websites, and on 162

occasion, issue notices of particularly noteworthy ENSO conditions. Other national 163

Page 6: El Niño-Southern Oscillation (ENSO) in a Changing Climate ... · 104 the “Climate Diagnostics Bulletin.” Initially, the forecast forum of the Bulletin featured 105 forecasts

6

meteorological services dedicate resources to ENSO forecast discussions that provide qualitative 164

and quantitative description of current conditions in the tropical Pacific, its likely upcoming state 165

(e.g., likelihood of El Niño or La Niña), and potential influences of ENSO on the regional 166

climate (e.g., rainfall and temperature). Many of these discussions are archived online, so users 167

can retrospectively examine previous forecast discussions to see what forecasters were 168

considering. Since 1997, the World Meteorological Organization has also put together a 169

discussion when ENSO conditions appear possible or have arrived. This discussion is updated 170

roughly four times per year and represents the international consensus of the member countries 171

that provide input to the ENSO discussion. 172

A more recent addition to operational ENSO outlooks is the implementation of ENSO 173

alert systems beginning between 2009-2015 (see table in L’Heureux et al. (2017) for an outline 174

of systems at CPC, BOM, and in Peru). Increasingly, forecasts quantify the probability of El 175

Niño, ENSO-neutral, and La Niña for upcoming seasons (this began in 2012 at CPC, in 176

conjunction with the IRI; L’Heureux et al., 2019). And with the advent of social media, 177

meteorological centers are increasingly leaning on these tools to alert users. The use of on-line 178

videos and blogs has also increasingly gained traction as a way to expand communication 179

platforms to those who would not otherwise read a technical ENSO discussion. 180

181

13.2. ENSO Predictability 182

183

What are the physical and dynamical mechanisms that allow ENSO to be predicted? 184

Conversely, which factors fundamentally limit ENSO forecast accuracy and skill? These are 185

questions about the predictability of ENSO. Their answers depend on the ENSO aspect of 186

interest, when the forecast is made, and the forecast lead time. Generally, ocean-based ENSO 187

indices are more predictable than atmospheric ones. ENSO predictability decreases with forecast 188

lead and is often lower for forecasts made during the boreal spring compared with the rest of 189

year, a characteristic known as the spring predictability barrier. Understanding ENSO 190

predictability is crucial for evaluating whether ENSO forecasts have the right level of 191

uncertainty. As forecast lead times increase, the range of possible outcomes becomes larger. If 192

the models that we use to predict ENSO have too large or too small of a range, our forecasts can 193

be underconfident or overconfident, respectively. 194

Page 7: El Niño-Southern Oscillation (ENSO) in a Changing Climate ... · 104 the “Climate Diagnostics Bulletin.” Initially, the forecast forum of the Bulletin featured 105 forecasts

7

Broadly speaking, two factors limit our ability to forecast ENSO. First, small errors in 195

our estimates of the current state of the ocean and the atmosphere (initial conditions) may grow 196

and lead to increasingly large forecast errors as we forecast further ahead. In ENSO forecasting, 197

the growth of initial condition errors is amplified by the air-sea interactions due to the strong 198

coupling in the tropics (Zebiak and Cane, 1987). When a forecast begins, the initial conditions 199

are our best guess of the current state of the atmosphere and ocean given the available 200

observations. However, observations themselves have measurement errors, not every quantity 201

can be observed, and the methods by which we include observations that have different spatial or 202

temporal resolution than the model (data assimilation) may introduce errors. Thus, the initial 203

conditions are imperfect. One way of accounting for the uncertainty in initial conditions is to 204

make multiple forecasts, or an ensemble, each starting from a slightly different initial condition 205

(Stockdale et al., 1998). The small differences between these initial conditions grow, and the 206

differences in the final forecasts provide an indication of the uncertainty due to initial condition 207

errors. These ensemble forecasts also contain representations of weather, which is chaotic and 208

unpredictable after a few weeks. Since these ensemble forecasts present a range of possible 209

outcomes, each of which is consistent with our best information, it is natural to express ENSO 210

forecasts using probabilities. The challenge for models is to produce the correct range of 211

possible outcomes given the initial conditions and unpredictability of the weather. 212

Second, all forecast models have important physical processes that are either missing or 213

represented imperfectly, often because they occur at smaller spatial and/or temporal scales than 214

the models are capable of resolving with the available computing capacity. The resulting “model 215

error” restricts how well models capture potentially realistic outcomes and therefore how well 216

they predict the likelihood of an ENSO event occurring. Current research aims to better 217

represent these processes by using higher resolution simulations, increased observations, and 218

improved parameterizations. Forecast uncertainty due to model error is harder to quantify than 219

that due to initial condition error since it pertains to “unknown unknowns,” but past forecast 220

performance can provide indications of model deficiencies and their impacts on forecasts. To the 221

extent that model error can be viewed as another source of uncertainty, it can be accounted for by 222

combining several different model ensembles into one larger “multi-model” ensemble (discussed 223

further in the next section). 224

Page 8: El Niño-Southern Oscillation (ENSO) in a Changing Climate ... · 104 the “Climate Diagnostics Bulletin.” Initially, the forecast forum of the Bulletin featured 105 forecasts

8

The presence or absence of various ocean conditions can be important for the extended 225

predictability of ENSO. From recharge oscillator theory, it is understood that equatorial warm 226

water volume (WWV; typically measured as the volume of water above 20°C between 5°S-5°N) 227

anomaly is an important predictor of future ENSO SSTs (Jin, 1997; Meinen and McPhaden, 228

2000). The WWV anomaly peaked 6-9 months before the SST peak during the period of 1980-229

1999 and 4-6 months from 2000-2010 (McPhaden, 2012). Therefore, WWV can act as an early 230

precursor of ENSO anomalies and, in particular, can provide some predictability beyond the 231

spring predictability barrier (Balmaseda et al., 1995; McPhaden, 2003; Newman et al., 232

2011). However, more recently it has been shown that WWV is a better predictor of La Niña 233

events than of El Niño events (Luo et al., 2008; Levine and McPhaden, 2016; Di Nezio et al., 234

2017; Santoso et al., 2017; Planton et al., 2018). Extratropical SST patterns can also impact the 235

generation of ENSO events. The North Pacific Meridional Mode (NPMM) has been shown to 236

lead ENSO events, particularly during the northern hemisphere winter and spring (Vimont et al., 237

2001; Vimont et al., 2003). Studies have suggested that the NPMM can either be forced from 238

extra-tropical winter storms and propagated equatorward via the wind-evaporation-SST feedback 239

(Vimont, 2010) or by the trade wind charging process (Anderson and Perez, 2015). Similarly, 240

equatorward propagation of anomalies in the southern hemisphere can also impact El Niño. The 241

South Pacific Meridional Mode (SPMM) can significantly impact the development of El Niño 242

events during the boreal summer and fall. Some studies have suggested that the SPMM can 243

influence the type of El Niño event that occurs during summer and fall months, the traditional 244

growth period of El Niño (You and Furtado, 2017), although the exact role of the SPMM still 245

remains unclear (Larson et al, 2018). 246

Other tropical regions also impact ENSO predictability (Cai et al., 2019). The tropical 247

Indian Ocean and the Indian Ocean Dipole interacts with ENSO and may contribute positively to 248

ENSO predictability at long lead times (Luo et al., 2010; Izumo et al., 2010). The Indian Ocean 249

has been suggested as a potential cause for the recent much anticipated non-El Niño event in 250

2014 and subsequent extreme event in 2015 (Dong and McPhaden, 2018). The tropical Atlantic, 251

including the “Atlantic Niño” and tropical North Atlantic SST anomaly may provide some 252

additional predictability for ENSO (Ham et al., 2013; Keenlyside et al., 2013; Luo et al., 2017), 253

but its role is modulated on multidecadal timescales (Martin-Rey et al., 2014). 254

Page 9: El Niño-Southern Oscillation (ENSO) in a Changing Climate ... · 104 the “Climate Diagnostics Bulletin.” Initially, the forecast forum of the Bulletin featured 105 forecasts

9

Atmospheric forcing in the tropical Pacific on weather timescales can have a significant 255

influence on ENSO development and evolution (Karspeck et al., 2006; Menkes et al., 2014). 256

Westerly wind bursts in the western and central Pacific act to force El Niño in two different 257

important ways. First, the westerly wind burst acts to push the warmer surface water in the 258

western Pacific towards the central Pacific (Anderson and McCreary, 1985; Puy et al., 2016; 259

Levine et al., 2017). This eastward extension of the warm pool appears important to the 260

formation of weaker, Central Pacific El Niño events (Ren and Jin 2013; Chen et al 2015). 261

Second, the westerly wind bursts act on the oceanic subsurface, displacing the thermocline and 262

increasing the WWV (McPhaden et al., 1998; Fedorov, 2002; Fedorov et al., 2015). This 263

thermocline displacement propagates eastward as a Kelvin wave, deepening the thermocline 264

across the entire Pacific over the course of 2-3 months and helping to create the incipient 265

conditions for stronger, Eastern Pacific El Niño events. Since westerly wind bursts are weather 266

time-scale events, they are inherently unpredictable on the timescale of seasonal forecasts. 267

However, given the nature of the interactions between westerly wind bursts and El Niño, 268

predicting the exact timing of the westerly wind burst is less essential than predicting that they 269

will occur over the course of the season (Roulston and Neelin, 2000; Levine and Jin, 2010). 270

Westerly wind bursts are modulated by the conditions in the tropical Pacific, for example, 271

occurring more frequently when the warm pool is extended eastward than when the warm pool is 272

in its normal position or has retreated westward (Lengaigne et al., 2004; Eisenman et al., 2005). 273

Although these aspects of westerly wind bursts reduce the deterministic predictability of ENSO, 274

it is important that they are correctly represented in the model in order to obtain reliable 275

probabilistic predictions of ENSO (Fedorov et al., 2003; Levine et al., 2016). 276

ENSO predictability also exhibits a pronounced seasonality, primarily due to two related 277

factors. First, over the last 50 years it appears that with one exception, all ENSO events peaked 278

during the boreal winter. Second, ENSO forecasts in both dynamical and statistical models are 279

generally much more skillful when they are started after boreal spring than before or during it 280

(Jin et al., 2008). A pronounced annual cycle in the eastern tropical Pacific, including the 281

development of the equatorial cold tongue from boreal summer to fall, drives seasonal variations 282

in the processes responsible for equatorial Pacific SST anomaly growth (Stein et al., 2010, 2015). 283

SST anomalies are most strongly damped during the boreal spring and most strongly amplified 284

during the boreal fall (Liu et al., 2019). Under these conditions, predictable signals grow weakly 285

Page 10: El Niño-Southern Oscillation (ENSO) in a Changing Climate ... · 104 the “Climate Diagnostics Bulletin.” Initially, the forecast forum of the Bulletin featured 105 forecasts

10

during the spring and can be overwhelmed by unpredictable factors. Also, all of the previous 286

factors that reduce the predictability of ENSO, challenges in observing the initial conditions or 287

the impact from the forcing, are potentially important for creating the spring predictability barrier 288

because the source of the SST anomaly is unimportant for its growth (Yu et al., 2012; Levine and 289

McPhaden, 2015). There has also been some suggestion that the seasonal cycle of the westerly 290

wind bursts impacts the seasonality of ENSO predictability, with the northern hemisphere spring 291

having the most westerly wind bursts, and the northern hemisphere summer and fall having the 292

least (Zheng and Zhu, 2010; Lopez and Kirtman, 2014). So, our lack of ability to predict ENSO 293

during the boreal spring is due to the time it takes the SST anomalies to grow and the likelihood 294

of the stochastic forcing occurring during specific seasons (Thomas et al., 2018). 295

296

13.3. ENSO Prediction Skill 297

298

The more closely a prediction corresponds to observations, the more skillful it is. 299

Forecast verification, the act of assessing skill, usually employs multiple skill measures, or 300

metrics. Choosing metrics is non-trivial and depends on user interests, including what aspects of 301

the forecast they want to evaluate. Skill can be evaluated in space or in time, it can assess a 302

single category or multiple categories, and it can be deterministic or probabilistic. Rigorous 303

verification also requires considering: (1) Start time: when is your forecast initiated? and (2) 304

Target time: what period are you forecasting? The forecast “Lead time” is then the time interval 305

between the start and target times. ENSO prediction skill is commonly displayed as a function 306

of some combination of start, target, or lead times. 307

ENSO skill is often verified using Niño indices, which measure SST anomalies averaged 308

within different regions along or near the equatorial Pacific Ocean. Probably the most common 309

SST index to measure skill is the Niño-3.4 index, which is based on SST anomalies in the east-310

central equatorial Pacific and is highly correlated with other variables and indicators of ENSO, 311

such as sea level pressure and rainfall over the tropical Pacific (Barnston et al., 1997; Trenberth, 312

1997), and with the extratropical response to ENSO (Kumar et al. 1996; L’Heureux et al., 2015). 313

Furthermore, unlike other variables within the tropical Pacific, a substantial investment has been 314

made to develop several homogenous, observationally-based SST climate records, which can be 315

used for skill validation over a long time period. So, while the Niño-3.4 index is not the only 316

Page 11: El Niño-Southern Oscillation (ENSO) in a Changing Climate ... · 104 the “Climate Diagnostics Bulletin.” Initially, the forecast forum of the Bulletin featured 105 forecasts

11

indicator of ENSO, it is commonly used to assess the quality of ENSO predictions, with many 317

model providers distributing real-time forecasts of Niño-3.4 (Barnston et al., 2012). That the 318

Niño-3.4 index is a single time series and has a near-Gaussian distribution also makes it 319

relatively easy to analyze. Also, tropical Pacific SST model forecasts are typically most skillful 320

in the Niño3.4 region (e.g., Jin et al., 2008; Newman and Sardeshmukh 2017). On the other hand, 321

not all ENSO events have the same SST pattern, so other Niño regions and indices also need to 322

be considered to entirely capture ENSO (Capotondi et al. 2015). 323

Figure 3 shows an evaluation of Niño-3.4 skill from the North American Multi-Model 324

Ensemble (based on the 1982-2018 record). Target times across the calendar year are on the x-325

axis, and the lead-time on the y-axis. The figure illustrates how far in advance the Niño-3.4 326

index can be predicted with skill above specified thresholds. The different colored bars 327

correspond to different correlation thresholds. So, if one desires at least 50% of Niño-3.4 index 328

variance to be predicted (r= 0.7), then ENSO can be predicted 5 months (for August-September) 329

to 10 months (for April-June) ahead, whereas Niño-3.4 skill in excess of r=0.9 (81% of variance 330

explained) is only possible at zero-month lead for June and July targets and up to 5 months in 331

advance for January targets. 332

Generally, the most trustworthy assessment of skill comes from assessing the skill of 333

predictions that are made in “real-time,” or on a regular, ongoing basis, as in an operational 334

framework. Barnston et al. (1994) noted that “performance in a real-time setting is the ultimate 335

test of the utility of a long-lead forecast.” One major advantage of verifying real-time predictions 336

is that forecast providers cannot pick only specific, possibly advantageous, forecast times that 337

could otherwise result in a biased verification. To truly assess the suitability of any model for 338

decision making and planning, it must consistently be initialized on the latest (previously 339

unknown) observations, run forward, and evaluated for its ability to capture ENSO evolution 340

over many forecasts. Because models are constructed with past observations already known and 341

based on previously observed physics or statistical relationships, the unrealized future is the only 342

truly independent scenario to objectively determine the skill of various models. On the other 343

hand, estimating skill from real-time forecasts can challenging because of the relatively short 344

record of real-time forecasts, as well as changing forecast methodologies. 345

In addition to real-time evaluation, another approach to assess ENSO skill is through 346

hindcasts, or reforecasts, when a model is run in forecast mode on known, past data. Akin to 347

Page 12: El Niño-Southern Oscillation (ENSO) in a Changing Climate ... · 104 the “Climate Diagnostics Bulletin.” Initially, the forecast forum of the Bulletin featured 105 forecasts

12

making a forecast run using the most recent observational data gathered within the past day or 348

so, the model is also run on the observed conditions of each day (or certain days) across many 349

years, going back several decades (usually to ~1980, when denser, satellite-based observations 350

that can initialize models first became available). In this fashion, forecast skill over a longer 351

period can be assessed across a variety of ENSO events, reducing (but not eliminating) 352

dependence upon sample size. 353

However, hindcast skill may differ from real-time predictions because of practical issues 354

involved with making real-time forecasts, such as missing or faulty data ingests and other 355

computing bugs (e.g., National Weather Service, 2016). Also, hindcast skill reflects the skill of 356

state-of the-art current generation forecast models, not the skill of previous generation models, 357

which are eventually phased out in real-time forecasting. This abandonment of older forecast 358

systems complicates determining whether operational ENSO skill has improved over time due to 359

advancements in forecast models and their assimilation systems. A long record of real-time 360

predictions, using the model systems in place at the time the prediction was made, would be 361

preferable to answer the question of whether model skill has improved over time. However, the 362

longest known real-time record is the IRI/CPC set of Niño-3.4 outlooks going back to February 363

2002 (International Research Institute for Climate and Society, 2002), indicates that natural, 364

decadal variations in skill cloaks any apparent change due to forecast technology (Barnston et al., 365

2012). For instance, despite the advancements in models through the 2000s, this decade was 366

accompanied by overall lower prediction skill compared to the 1990s (Xue et al., 2013). 367

While the first seasonal ENSO predictions were made in the early 1980s, prediction skill 368

was not evaluated and compared across models until the 1990s (Barnston et al., 1994). Three 369

broad types of models have been used to predict ENSO: (1) Statistical, (2) Hybrid, and (3) 370

Dynamical models. Statistical models are based on historical relationships in the observational 371

record and use analysis techniques such as linear regression, neural networks or other machine 372

learning methods. Dynamical models are based on physical equations and parameterizations of 373

the coupled ocean-atmosphere climate system and typically require supercomputing resources. 374

Hybrid models use a dynamical ocean model and infer the atmosphere from statistical 375

relationships. Hybrid models are not presently used in ENSO forecasting, though they played a 376

role in the past (Latif et al., 1998). Within the operational centers, dynamical model 377

development has been emphasized since the 1990s, and little statistical model development has 378

Page 13: El Niño-Southern Oscillation (ENSO) in a Changing Climate ... · 104 the “Climate Diagnostics Bulletin.” Initially, the forecast forum of the Bulletin featured 105 forecasts

13

taken place since the 2000s. The benefits of this focus may have been realized with Tippett et al. 379

(2012), Barnston et al. (2012), and van Oldenborgh (2005a,b) showing that dynamical models 380

have very slightly outperformed their statistical counterparts for the Niño indices during the 381

previous decade, although the differences for Niño3.4 are small and its statistical significance is 382

debatable (Newman and Sardeshmukh 2017). 383

One of the most notable features in ENSO skill, which continues to bedevil forecasters, is 384

a relatively lower skill level for forecasts made through the boreal spring. The previous section 385

discusses ideas for the springtime prediction or predictability barrier. Despite years of model 386

improvement, its endurance suggests real, fundamental intrinsic limits on skill. This minimum in 387

forecast skill, which impacts target months during the summer and early fall is also seen in 388

Figure 3. Forecasts made in the spring suffer a subsequent reduction in skill that is most 389

noticeable up to five months later (e.g., a model initialized in March or April is associated with a 390

visible reduction of skill for August-September targets). 391

Despite the decrease in skill across the spring, state-of-the-art dynamical models tend to 392

outperform many statistical models owing to their higher skill for spring and summer targets 393

(van Oldenborgh, 2005a; Barnston et al. 2012; Barnston et al., 2017). Among other possibilities, 394

this performance may be due to the advanced assimilation systems of dynamical models, which 395

more rapidly ingest the latest observations, such as a recent westerly wind burst, which can be a 396

precursor for El Niño. Most statistical models tend to be coarser, using monthly or seasonal 397

averaged data, so do not resolve these short-term developments (Barnston et al. 1999). The 398

relatively short and uncertain observational record may also limit statistical models. When 399

trained on multi-centennial output from climate simulations, some statistical models have skill 400

comparable to dynamical model skill (e.g., Chen et al. 2016; Ding et al. 2018). Regardless, 401

statistical models play a valuable role in providing benchmarks for comparison, and their lower 402

dimensionality can help provide insight on important mechanisms and sources of predictability. 403

Further, statistical methods are often applied to dynamical model output to calibrate and combine 404

models (e.g., Stephenson et al., 2005). 405

A hindcast record of sufficient length allows for the identification of robust physical 406

model errors in both forecast mean and uncertainty. This allows the skill of dynamical models to 407

be considerably enhanced by bias correction, a forecast anomaly is developed by removing the 408

lead-time dependent model climatology from the total fields (Stockdale, 1997; van Oldenborgh, 409

Page 14: El Niño-Southern Oscillation (ENSO) in a Changing Climate ... · 104 the “Climate Diagnostics Bulletin.” Initially, the forecast forum of the Bulletin featured 105 forecasts

14

2005b). Not only does this process remove mean biases, but also it removes any systematic 410

errors in the seasonal cycle, as well as lead-time dependent drift of the model climatology away 411

from the observed one. In this sense, all dynamical model forecasts are “statistically corrected.” 412

Statistical methods are occasionally also used to estimate flux correction terms within the model 413

integration itself in order to reduce certain climate model errors, such as cold tongue and double 414

ITCZ biases (Magnusson et al., 2013). 415

Increased skill from dynamical models is partly due to the introduction of multiple 416

member ensembles, which capture the range of uncertainty emanating from imperfect initial 417

conditions and the unpredictable evolution of the atmosphere. More widespread availability of 418

supercomputing resources has also allowed for combinations of ensembles associated with 419

multiple models that are each run (and bias corrected) in a similar configuration, such as the 420

North American Multi-Model Ensemble (NMME; Kirtman et al., 2014), European Multi-Model 421

Seasonal-to-Interannual Prediction project (EUROSIP; Palmer et al., 2004), and Asia-Pacific 422

Economic Cooperation (APEC) Multi-Model forecasts (Min et al., 2009). The use of multiple 423

model ensembles generally improves skill over a single model (Kirtman et al., 2001; Jin et al., 424

2008; Weisheimer et al. 2009; DelSole et al., 2014) and, perhaps more importantly, estimates of 425

forecast uncertainty. Multi-model systems, where each model is accompanied by a long 426

reforecast, improve the statistical correction of forecasts not only about the mean, but also in 427

their variance. Hence, their benefit is realized through improved reliability scores and other 428

types of probabilistic verification metrics (Tippett and Barnston, 2008; Barnston et al., 2015; 429

Tippett et al., 2017). However, ENSO variability errors common to all models (e.g., a westward 430

displacement of the SST anomaly; Newman and Sardeshmukh 2017) are not as easily corrected 431

(Ding et al. 2018). Recent efforts have been made to increase the ocean resolution and better 432

parameterize tropical convection, which appears to reduce bias and improve variability of 433

various aspects of the coupled ocean-atmosphere system (MacLachlan et al. 2015; Stockdale et 434

al., 2018; Johnson et al., 2019). 435

In general, the prediction skill for other SST regions of the tropical Pacific is lower than 436

that for Niño-3.4 and the east-central Pacific (Luo et al., 2008; Magnusson et al., 2013; Newman 437

and Sardeshmukh 2017), which has implications for forecasting the continuum of SST types, or 438

ENSO flavors. Central Pacific events tend to have SST anomalies that are relatively larger 439

within the Niño-4 region, while Eastern Pacific events have largest amplitude closer to South 440

Page 15: El Niño-Southern Oscillation (ENSO) in a Changing Climate ... · 104 the “Climate Diagnostics Bulletin.” Initially, the forecast forum of the Bulletin featured 105 forecasts

15

America, within the Niño-3 and Niño-1+2 regions. Weaker El Niños often coincide with Central 441

Pacific events (Fig. 3 in Capotondi et al., 2015; Timmerman et al., 2018), which can be trickier 442

to predict because of their lower amplitudes and generally later onset times (Imada et al., 2015). 443

For example, Figure 4 shows the February 2019 prediction from the NMME where the ensemble 444

mean lies near El Niño thresholds for all lead times, but the spread of ensemble members 445

encompasses conditions ranging from La Niña to El Niño. Thus, in an operational setting, it is 446

easy to be more confident of an El Niño forecast when most ensembles are well above required 447

event thresholds (+0.5°C or +0.8°C) than of a weaker event when the forecast uncertainty 448

significantly spans other outcomes. 449

While fewer in number, predictions related to the strongest El Niños, or Eastern Pacific 450

events, tend to underestimate peak intensity for most lead times beyond a couple months. This 451

was an issue for real time predictions of both the 1997-98 El Niño (Barnston et al., 1999) and the 452

2015-16 El Niño (L’Heureux et al., 2017). However, these stronger events were also more 453

skillfully predicted further in advance (e.g., 6-9 months) of their boreal winter peaks, in part due 454

to persistent and strong signals that grew through the preceding year (Luo et al., 2016; Zheng and 455

Yu, 2017). In recent years, there have been growing efforts to better resolve the different spatial 456

flavors of ENSO in prediction models because of its potential importance on regional climate 457

variability (Hendon et al. 2009). For example, the new operational forecast model from BOM, 458

the ACCESS-s1, appears to better distinguish different flavors compared to its predecessor 459

model (Hudson et al., 2017). 460

461

13.4. Decadal Variation in ENSO and its Skill 462

463

Both ENSO forecast skill and predictability have varied substantially over the past few 464

decades (e.g., Balmeseda et al., 1995; Barnston et al., 2012; Newman and Sardeshmukh, 2017; 465

Huang et al., 2017; Ding et al., 2019). This is illustrated in Figure 5, which uses a measure called 466

pattern correlation to show how well the ensemble-mean SST forecast for a six-month lead time 467

matches observations within the ENSO region spanning the central and eastern tropical Pacific. 468

The resulting forecast skill, smoothed to emphasize the year-to-year variations, is shown for an 469

empirical-dynamical model called the Linear Inverse Model (LIM; Newman and Sardeshmukh, 470

2017), three physical-dynamical ensembles (the NMME operational multi-model ensemble, the 471

Page 16: El Niño-Southern Oscillation (ENSO) in a Changing Climate ... · 104 the “Climate Diagnostics Bulletin.” Initially, the forecast forum of the Bulletin featured 105 forecasts

16

ECMWF SEAS5 25-member single model ensemble, and the SINTEX-F 9-member single model 472

ensemble; Luo et al., 2008), and a statistical approach (model-analogs; Ding et al., 2018; 2019) 473

that mimics ENSO forecasts made by traditional multi-model ensembles by finding matches to 474

observed oceanic surface anomalies within two different sets of long climate model simulations 475

from either NMME or CMIP5 models. The LIM hindcasts are determined using a ten-fold cross-476

validation, and the NMME, SEAS5, and SINTEX-F hindcasts have been bias-corrected 477

separately by model; model-analogs are bias-corrected by construction. Also shown is the ENSO 478

predictability, or average expected skill, estimated from the LIM as a function of the “signal-to-479

noise” ratio between its predicted ENSO amplitude and its forecast uncertainty (or ensemble 480

spread), assumed to depend only upon forecast lead time. 481

Two things are immediately clear from this figure. First, despite using fairly different 482

modeling techniques, all three approaches (models, LIM, analogs) often have similar year-to-483

year skill variations. Note that the LIM expected skill has similar variations, showing how 484

predictability estimates can be used to diagnose the potential for actual forecast skill. For this 485

definition of skill, it might be anticipated that the model forecasts have greater expected and 486

actual skill at times of larger forecast amplitudes, and this can be seen to be true for forecast skill 487

values during years of moderate to strong ENSO events (indicated by shading in the figure). 488

Additionally, since the expected skill can be determined at the time of forecast issuance, it could 489

be used to indicate when confidence in the spatial structure of the forecast anomaly is expected 490

to be relatively high. 491

Second, while no long-term trend in ENSO skill is apparent since 1961 (Ding et al., 492

2019), there are some extended periods when forecasts appear generally more skillful than 493

during other periods. Recently, the relatively low skill that existed for much of the first decade of 494

the 2000’s led some researchers to suggest that ENSO predictability had been reduced, possibly 495

because of some fundamental change to average tropical conditions (e.g., Barnston et al., 2012; 496

Zhao et al., 2016). Moreover, some studies have posited that the nature of ENSO itself has 497

changed over the last few decades, with an increased prevalence of Central Pacific events 498

possibly also due to changes in the average ocean, or base state, conditions (Lee and McPhaden, 499

2010), including those potentially driven by climate change (Yeh et al., 2009). For example, it 500

has been suggested that WWV and related thermocline feedbacks had weaker impacts on ENSO 501

SST anomaly development during the 2000’s, causing more Central Pacific events with less 502

Page 17: El Niño-Southern Oscillation (ENSO) in a Changing Climate ... · 104 the “Climate Diagnostics Bulletin.” Initially, the forecast forum of the Bulletin featured 105 forecasts

17

predictability (McPhaden, 2012; Neske and McGregor, 2018). Such a base state change on long 503

time scales, if predictable, could lead to long-term ENSO forecasts of decadal ENSO variability. 504

On the other hand, some researchers have suggested that variations in both the amplitude 505

and type of ENSO events are largely driven by noise -- that is, by random and essentially 506

unpredictable year-to-year weather variations in both the tropics and extratropics (e.g., Newman 507

et al., 2011; Thomas et al., 2018). In this case, year-to-year variations in ENSO forecast skill 508

might also occur simply due to differing noise events, and apparent base state changes might be 509

the residual of random ENSO events rather than a forcing of them (Flügel et al., 2004; 510

Wittenberg et al., 2014; Kumar et al., 2015; Lee et al., 2016; Newman and Sardeshmukh, 2017). 511

Then, ENSO might not be predictable for forecast leads extending much beyond its life cycle. 512

For example, since the LIM is constructed so that its predictable dynamics are unchanging over 513

the 1961-2010 period (apart from cross-validation, which has minor impact), its expected 514

decadal variations of ENSO activity are solely a consequence of unpredictable decadal variations 515

in weather noise. Likewise, the LIM expects seasonal forecast skill to increase whenever, by 516

chance, a substantial ENSO precursor is initiated. These expectations (dotted line in Fig. 5) were 517

in fact largely realized by all the hindcasts in Figure 5, with periods of enhanced skill in the mid 518

1970’s and late 1990’s, coinciding with enhanced ENSO activity, and reduced skill in the late 519

1970’s/early 1980’s and early 2000’s, coinciding with reduced ENSO activity. 520

Of course, decadal variations in ENSO events and predictability could result from both 521

weather noise and base state oceanic changes (e.g., Aiken et al., 2013; Capotondi and 522

Sardeshmukh, 2017), as well as the convolution between them (e.g. Levine et al., 2016), with 523

consequent impacts on forecast skill. Still, if changes to the base state and its related ENSO 524

variability are predictable on decadal timescales, current physical models seem unable to capture 525

them. At present, on average these models have little SST skill within the ENSO region for 526

forecast leads greater than about two years (Newman, 2013). Some recent studies, however, have 527

suggested that longer lead forecast skill may exist for a few isolated events (Gonzalez and 528

Goddard, 2015), particularly some La Niñas that either occur as a result of a transition from a 529

strong El Niño (e.g., Thomas et al., 2018) or that follow a previous La Niña as part of a 2-yr La 530

Niña event (DiNezio et al., 2017). 531

532

13.5. Recent ENSO Prediction Challenges 533

Page 18: El Niño-Southern Oscillation (ENSO) in a Changing Climate ... · 104 the “Climate Diagnostics Bulletin.” Initially, the forecast forum of the Bulletin featured 105 forecasts

18

534

Some current challenges in ENSO prediction were highlighted and illustrated by the 535

surprising evolution of ENSO during 2014 when early indications that a strong El Niño could be 536

brewing (Carrington et al., 2014) turned out to be a false alarm. Despite a strong downwelling 537

oceanic Kelvin wave in the boreal spring of 2014 and climate model forecasts for El Niño, only 538

marginally warm tropical Pacific conditions with little atmospheric coupling were observed 539

during the boreal winter 2014-15 (McPhaden, 2015). Better understanding of the reasons for 540

such problematic ENSO forecasts can identify obstacles to improved ENSO prediction. Many 541

obstacles are related to the physics of the climate system, but better quantifying and 542

communicating uncertainty is also a challenge since some forecasting systems, such as the 100-543

member North American multi-model ensemble, suggested the observed evolution was unlikely 544

but still possible (Figure 6, top), a view supported by the EUROSIP calibrated multi-model 545

forecast (Figure 6, bottom). 546

Another challenge is forecasting surface winds over the equatorial Pacific Ocean. These 547

winds can have a considerable impact on the evolution and strength of ENSO (Menkes et al., 548

2014; Hu and Federov, 2016; Levine and McPhaden, 2016; Takahashi and Dewitte, 2016), and 549

the unpredicted and sudden onset of easterly wind anomalies during July 2014 was one of the 550

reasons for reduced oceanic warmth during the rest of the year. Unfortunately, prediction errors 551

in tropical winds arise quickly and are large by day 5 (Martin et al., 2010). However, some 552

aspects of these winds may arise from low-frequency SST variability and may be more 553

predictable (Eisenman et al., 2005; Tziperman and Yu, 2007; Levine and Jin, 2017; Capotondi et 554

al., 2018; Ineson et al., 2018). The extent to which ENSO forecast uncertainty is due to 555

limitations in forecasting the surface winds over the equatorial Pacific Ocean remains an open 556

question. 557

The extent of the role of other ocean basins and the off-equatorial Pacific Ocean also 558

needs to be clarified. For instance, conditions in 2014 across the Indian Ocean (Dong and 559

McPhaden, 2018) and North and South Pacific Oceans (Larson and Kirtman, 2015; Min et al., 560

2015; Zhu et al., 2016; Wang and Hendon, 2017) have been labeled as possible culprits for the 561

aborted El Niño. Non-equatorial Pacific anomalies can be quite critical for the development of 562

ENSO in general (Vimont et al, 2001; Chang et al., 2007; Di Lorenzo et al., 2010; Wang et al., 563

2012; Boschat et al., 2013; Keenlyside et al., 2013; Zhang et al., 2014; Luo et al., 2017; Pegion 564

Page 19: El Niño-Southern Oscillation (ENSO) in a Changing Climate ... · 104 the “Climate Diagnostics Bulletin.” Initially, the forecast forum of the Bulletin featured 105 forecasts

19

and Selman, 2017; You and Furtado, 2017), but it is another task to translate this knowledge into 565

improved predictions (Larson and Kirtman, 2014; Larson et al., 2018). So, the question remains 566

whether current climate models appropriately capture and predict these extratropical-tropical 567

linkages. 568

While 2014-15 was a recent example of a challenging ENSO forecast; prediction and 569

characterization of marginal El Niño events are always particularly difficult. Since ENSO is a 570

complex, coupled ocean-atmosphere system that is measured using multiple regions and 571

variables, those indicators are not necessarily all in alignment during borderline events. In 2014 572

some national meteorological services issued watches or alerts, indicating that conditions were 573

favorable for the onset of El Niño later in the year. From the standpoint of SSTs in the Niño-3.4 574

region, conditions later in the year were consistent with El Niño, but several forecast centers 575

(e.g., NOAA, BOM) never declared the onset of El Niño due to the lack of significant 576

atmospheric coupling over the equatorial Pacific (L’Heureux, 2015; Santoso et al., 2018). 577

Therefore, while a major emphasis of real-time model forecasts is on SSTs, there is no guarantee 578

of a concomitant response in other indicators that may be important for impacts. Currently, 579

operational predictions of atmospheric-based ENSO indicators, such as the Southern Oscillation 580

Index (SOI) and Outgoing Longwave Radiation (OLR) are not commonplace and are likely of 581

lower skill than SST indices. Figure 7 shows log skill scores from the Climate Forecast System 582

(version 2) for three different atmosphere-based ENSO indices, in addition to the Niño-3.4 SST 583

index. In this particular model, predictions of the Niño-3.4 SST index are clearly more skillful 584

than the atmospheric indices, though the Equatorial SOI appears to provide skill beyond the two 585

OLR based indices. 586

Finally, with the planet warming in association with anthropogenic climate change, SST 587

values have begun to include warming trends not necessarily related to ENSO (Solomon and 588

Newman 2012). For example, the 2015-16 El Niño appears to be record breaking based on 589

Niño-3.4 SST anomalies, but if detrended then it was not as warm and suggests non-linear 590

feedbacks were not triggered in the eastern Pacific to the same extent as previous El Niño events 591

(Santoso et al. 2017; L’Heureux et al., 2017). Newman et al. (2018) pointed out that SST 592

anomalies in the westernmost Niño-4 region were unprecedented in the observational record and 593

unlikely to have occurred without anthropogenic warming. However, it is not clear if or how this 594

warming is associated with a change in ENSO variability itself, which is an important 595

Page 20: El Niño-Southern Oscillation (ENSO) in a Changing Climate ... · 104 the “Climate Diagnostics Bulletin.” Initially, the forecast forum of the Bulletin featured 105 forecasts

20

consideration given continued uncertainty in the projections of future El Niño variability in the 596

eastern Pacific Ocean (e.g., Stevenson et al. 2017; Cai et al., 2018). Thus, an emerging challenge 597

in real-time forecasting is to untangle El Niño dynamics from long-term climate trends. 598

599

13.6. Concluding Remarks 600

601

Our understanding of ENSO has matured, and with it forecast systems have improved in 602

their representation of nature and handling of observations. Now skillful predictions of ENSO 603

are routinely made by operational centers around the world. The scientific and institutional 604

investments have improved ENSO forecasting have also enabled seasonal predictions of other 605

climate anomalies, which provide advance guidance to a wide range of decision makers and 606

users. While the infrastructure to predict ENSO and communicate its impacts has significantly 607

advanced since the 1980s, the community now has greater appreciation for variation in skill over 608

the span of decades and tremendous diversity of ENSO types, which are not equally well 609

predicted. 610

This chapter has focused on the history of ENSO prediction, both its skill and its potential 611

predictability, and highlighted some continuing prediction challenges that were illustrated by 612

2014-16 evolution of ENSO. It is clear that the upcoming decades will challenge the prediction 613

communities as anthropogenic climate change becomes more prominent even while its impact on 614

ENSO remain debated. Significant changes in the nature of ENSO, such as in its amplitude or 615

frequency, would almost certainly impact our ability to predict it, and in ways we may not 616

expect. Climate change could also have a substantial impact on the tools that forecasters rely 617

upon, which critically assume a long (~30 year) stationary record, whether as a training period 618

for statistical models or to bias correct dynamical models. Significant changes in the statistics of 619

ENSO, not accounted for in these past records, would greatly increase the uncertainty of ENSO 620

predictions. Fortunately, or unfortunately, forecasters and users making decisions off these 621

forecasts will lead the pack in experiencing the consequences of ENSO in a warming climate. 622

623

Acknowledgements 624

625

Page 21: El Niño-Southern Oscillation (ENSO) in a Changing Climate ... · 104 the “Climate Diagnostics Bulletin.” Initially, the forecast forum of the Bulletin featured 105 forecasts

21

We especially thank these individuals for their time in helping us with the History of ENSO 626

section: Tony Barnston (NOAA/IRI, retired), Grant Beard (BOM, retired), Mike Halpert 627

(NOAA), Vernon Kousky (NOAA, retired), Neville Nicholls (formerly BOM, currently Monash 628

University), Robert Reeves (NOAA, retired), and Chet Ropelewski (NOAA/IRI, retired). 629

630

Figure Captions: 631

632

Figure 1. A timeline of events in ENSO forecasting. 633

634

Figure 2. The first ENSO Diagnostics Advisory issued at NOAA Climate Prediction Center in 635

March 1986, which assessed an emerging warm episode. Photo Credit: Robert Reeves. 636

637

Figure 3. The forecast lead-time (y-axis) when Niño-3.4 index skill exceeds certain correlation 638

thresholds. On the x-axis, the target month of the forecast is presented. Model data is based on 639

the ensemble average of ~100 members from the North American Multi-Model Ensemble from 640

1982 to 2018. Departures in the Niño-3.4 index are based on 1982-2010 monthly averages. 641

642

Figure 4. Predictions of the Niño-3.4 index from the North American Multi-Model Ensemble 643

initialized at the beginning of February 2019. Predictions are shown out to 12 months lead-time 644

or through January 2020. The black solid line are the observations based on daily OISST, and 645

the dashed line is the ensemble average. Remaining lines show the individual members. 646

Departures in the Niño-3.4 index are based on 1982-2010 monthly averages. 647

648

Figure 5. Year-to-year variations of model hindcast/forecast skill in the tropical Pacific ENSO 649

region for monthly forecasts initialized during 1961-2015, for six month forecast lead time. The 650

measure of skill used is the pattern correlation between ensemble mean forecast and 651

corresponding observed monthly SST anomalies within the region bounded by 20ºS-20ºN, 652

170ºE-70ºW, smoothed with a 13-month running mean filter to emphasize annual variations. All 653

anomalies are determined relative to a 1961-2015 monthly climatology. The light red (blue) 654 shaded epochs indicate periods of >1 sigma warm (< −1 sigma cold) Niño3.4 seasonal SST 655

anomalies. Skill is shown for the Linear Inverse Model (LIM), the North American Multi-Model 656

Page 22: El Niño-Southern Oscillation (ENSO) in a Changing Climate ... · 104 the “Climate Diagnostics Bulletin.” Initially, the forecast forum of the Bulletin featured 105 forecasts

22

Ensemble (NMME) operational ensemble, the ECMWF SEAS5 ensemble, the JAMSTEC 657

SINTEX-F ensemble, and model-analogs employed on either a 4-model subset of the NMME 658

models or a 10-model subset of the CMIP5 models. The mean expected skill (a predictability 659

estimate) from the LIM is also shown. The LIM is trained from tropical sea surface analyses over 660

the 1958-2011 period, with hindcasts determined by a ten-fold cross validation technique. The 661

NMME and SEAS5 hindcasts are bias corrected over the 1982-2009 period. Therefore, all 662

forecasts after 2011 for the LIM, NMME, SEAS5, and SINTEX-F models are independent of 663

observations during that period. There is no training period for the model-analog forecasts, since 664

they are based on analogs to tropical IndoPacific sea surface observations drawn from 665

independent multi-centennial unitialized model simulations. 666

667

Figure 6. (top panel) Predictions of the Niño-3.4 index from the North American Multi-Model 668

Ensemble initialized at the beginning of April 2014. Predictions are shown out to 12 months 669

lead-time or through March 2015. The black solid line are the observations based on daily 670

OISST, and the dashed line is the ensemble average. Remaining lines show the individual 671

members. Departures in the Niño-3.4 index are based on 1982-2010 monthly averages. (bottom 672

panel) Predictions of the Niño-3.4 index from the EUROSIP multi-model system initialized on 673

the 1st April 2019, according to the operational calibrated product. Various percentiles of the 674

distribution are shown, based on a multi-model combination of variance-corrected ensemble 675

forecasts, with an additional calibration of the forecast uncertainty to match historic forecast 676

error. 677

678

Figure 7. Log skill score (LSS) from hindcasts (1981-2010) of the NCEP Climate Forecast 679

System (CFS) version 2. Skill is presented as a function of target month (x-axis) and lead-time 680

(y-axis). The top left panel shows the skill of the Niño-3.4 SST index region (5°S-5°N, 170°W-681

120°W), the bottom left panel shows the skill of the Equatorial Southern Oscillation Index 682

(surface pressure differences based 5°S-5°N, 80°W-130°W minus 5°S-5°N, 90°E-140°E), the top 683

right panel shows the skill of the Eastern Pacific Outgoing Longwave Radiation (OLR) index 684

(5°S-5°N, 160°W-110°W; Chiodi and Harrison, 2013), and the bottom right panel shows the 685

skill of the Central Pacific OLR index (5°S-5°N, 170°E-140°W; L’Heureux et al., 2015). 686

Page 23: El Niño-Southern Oscillation (ENSO) in a Changing Climate ... · 104 the “Climate Diagnostics Bulletin.” Initially, the forecast forum of the Bulletin featured 105 forecasts

23

Positive values of LSS show forecasts that are better than a climatological forecast and negative 687

values of LSS show where forecasts are worse than climatology. 688

689

Page 24: El Niño-Southern Oscillation (ENSO) in a Changing Climate ... · 104 the “Climate Diagnostics Bulletin.” Initially, the forecast forum of the Bulletin featured 105 forecasts

manuscript submitted to AGU Books

References

Aiken, C. M., Santoso, A., McGregor, S., & England, M. H. (2013, 2018/11/14).The 1970’s shift in ENSO dynamics: A linear inverse model perspective. Geo-phys. Res. Lett., 40 (8), 1612–1617. doi: 10.1002/grl.50264

Anderson, D. L. T., & McCreary, J. P. (1985). Slowly Propagating Disturbances ina Coupled Ocean-Atmosphere Model. J. Atmos. Sci., 42 (6), 615-630. doi: 10.1175/1520-0469(1985)042h0615:SPDIACi2.0.CO;2

Balmaseda, M. A., Davey, M. K., & Anderson, D. L. T. (1995). Decadal and Sea-sonal Dependence of ENSO Prediction Skill. J. Climate, 8 (11), 2705-2715. doi:10.1175/1520-0442(1995)008h2705:DASDOEi2.0.CO;2

BARNETT, T., GRAHAM, N., CANE, M., ZEBIAK, S., DOLAN, S., O’BRIEN,J., & LEGLER, D. (1988). On the Prediction of the El Nino of 1986-1987.Science, 241 (4862), 192–196. doi: 10.1126/science.241.4862.192

Barnston, A. G., Chelliah, M., & Goldenberg, S. B. (1997). Documentation ofa highly ENSO-related SST region in the equatorial pacific: Research note.Atmos.-Ocean, 35 (3), 367-383. doi: 10.1080/07055900.1997.9649597

Barnston, A. G., Glantz, M. H., & He, Y. (1999). Predictive skill of statistical anddynamical climate models in sst forecasts during the 1997-98 el nino episodeand the 1998 la nina onset. Bull. Amer. Meteor. Soc., 80 (2), 217-244. doi:10.1175/1520-0477(1999)080h0217:PSOSADi2.0.CO;2

Barnston, A. G., Tippett, M. K., L’Heureux, M. L., Li, S., & DeWitt, D. G. (2012).Skill of real-time seasonal ENSO model predictions during 2002-11: Is ourcapability increasing? Bull. Amer. Meteor. Soc., 93 (5), 631-651. doi:10.1175/BAMS-D-11-00111.1

Barnston, A. G., Tippett, M. K., Ranganathan, M., & L’Heureux, M. L. (2017).Deterministic skill of enso predictions from the north american multimodelensemble. Climate Dyn., 1–20. doi: 10.1007/s00382-017-3603-3

Barnston, A. G., Tippett, M. K., van den Dool, H. M., & Unger, D. A. (2015). To-ward an Improved Multimodel ENSO Prediction. J. Appl. Meteor. Climatol.,54 (7), 1579-1595. doi: 10.1175/JAMC-D-14-0188.1

Barnston, A. G., van den Dool, H. M., Zebiak, S. E., Barnett, T. P., Ji, M., Ro-denhuis, D. R., . . . Livezey, R. E. (1994). Long-Lead Seasonal Forecasts-Where Do We Stand? Bull. Amer. Meteor. Soc., 75 (11), 2097-2114. doi:10.1175/1520-0477(1994)075h2097:LLSFDWi2.0.CO;2

Battisti, D. S., & Sarachik, E. S. (1995). Understanding and predicting ENSO. Rev.Geophys., 33 (S2), 1367-1376. doi: 10.1029/95RG00933

Boschat, G., Terray, P., & Masson, S. (2013). Extratropical forcing of ENSO. Geo-phys. Res. Lett., 40 (8), 1605-1611. doi: 10.1002/grl.50229

Cai, W., Wang, G., Dewitte, B., Wu, L., Santoso, A., Takahashi, K., . . . McPhaden,M. J. (2018). Increased variability of eastern pacific el nino under greenhousewarming. Nature, 564 (7735), 201–206. doi: 10.1038/s41586-018-0776-9

Cai, W., Wu, L., Lengaigne, M., Li, T., McGregor, S., Kug, J.-S., . . . Chang,P. (2019). Pantropical climate interactions. Science, 363 (6430). doi:10.1126/science.aav4236

Cane, M., & Sarachik, E. (1991). Prospectus: A TOGA Program on Seasonal to In-terannual Prediction. UCAR Productions, NOAA Climate and Global ChangeProgram Special Report, no. 4 , 1-46.

Cane, M. A., Zebiak, S. E., & Dolan, S. C. (1986, 06 26). Experimental forecasts ofel nino. Nature, 321 , 827 EP -. Retrieved from https://doi.org/10.1038/321827a0

Capotondi, A., & Sardeshmukh, P. D. (2017, 2018/11/14). Is El Nino really chang-ing? Geophys. Res. Lett., 44 (16), 8548–8556. doi: 10.1002/2017GL074515

Capotondi, A., Sardeshmukh, P. D., & Ricciardulli, L. (2018). The Nature of theStochastic Wind Forcing of ENSO. J. Climate, 31 (19), 8081-8099. doi: 10.1175/JCLI-D-17-0842.1

–2–

Page 25: El Niño-Southern Oscillation (ENSO) in a Changing Climate ... · 104 the “Climate Diagnostics Bulletin.” Initially, the forecast forum of the Bulletin featured 105 forecasts

manuscript submitted to AGU Books

Capotondi, A., Wittenberg, A. T., Newman, M., Lorenzo, E. D., Yu, J.-Y., Bracon-not, P., . . . Yeh, S.-W. (2015). Understanding ENSO diversity. Bull. Amer.Meteor. Soc., 96 (6), 921-938. doi: 10.1175/BAMS-D-13-00117.1

Carrington, D., Godenberg, S., & Redfearn, G. (2014). How El Nino will change theworld’s weather in 2014. The Guardian.

Chang, P., Zhang, L., Saravanan, R., Vimont, D. J., Chiang, J. C. H., Ji, L., . . .Tippett, M. K. (2007). Pacific meridional mode and El Nino-Southern Oscilla-tion. Geophys. Res. Lett., 34 (16). doi: 10.1029/2007GL030302

Chen, C., Cane, M. A., Henderson, N., Lee, D. E., Chapman, D., Kondrashov, D.,& Chekroun, M. D. (2016). Diversity, Nonlinearity, Seasonality, and MemoryE↵ect in ENSO Simulation and Prediction Using Empirical Model Reduction.J. Climate, 29 (5), 1809-1830. doi: 10.1175/JCLI-D-15-0372.1

Chen, D., Lian, T., Fu, C., Cane, M. A., Tang, Y., Murtugudde, R., . . . Zhou, L.(2015, 04 13). Strong influence of westerly wind bursts on El Nino diversity.Nature Geosci , 8 , 339 EP -.

DelSole, T., Nattala, J., & Tippett, M. K. (2014). Skill improvement from increasedensemble size and model diversity. Geophys. Res. Lett., 41 (20), 7331-7342. doi:10.1002/2014GL060133

Di Lorenzo, E., Cobb, K. M., Furtado, J. C., Schneider, N., Anderson, B. T.,Bracco, A., . . . Vimont, D. J. (2010, 10 17). Central Pacific El Nino anddecadal climate change in the North Pacific Ocean. Nature Geosci , 3 , 762 EP-.

DiNezio, P. N., Deser, C., Okumura, Y., & Karspeck, A. (2017). Predictabilityof 2-year La Nina events in a coupled general circulation model. Climate Dyn.,49 (11), 4237–4261. doi: 10.1007/s00382-017-3575-3

Ding, H., Newman, M., Alexander, M. A., & Wittenberg, A. T. (2018). SkillfulClimate Forecasts of the Tropical Indo-Pacific Ocean Using Model-Analogs. J.Climate, 31 (14), 5437-5459. doi: 10.1175/JCLI-D-17-0661.1

Dong, L., & McPhaden, M. J. (2018). Unusually warm Indian Ocean sea surfacetemperatures help to arrest development of El Nino in 2014. Sci. Rep., 8 (1),2249. doi: 10.1038/s41598-018-20294-4

Eisenman, I., Yu, L., & Tziperman, E. (2005). Westerly Wind Bursts: ENSO’s TailRather than the Dog? J. Climate, 18 (24), 5224-5238. doi: 10.1175/JCLI3588.1

Fedorov, A. V. (2002). The response of the coupled tropical ocean?atmosphere towesterly wind bursts. Quart. J. Roy. Meteor. Soc., 128 (579), 1-23. doi: 10.1002/qj.200212857901

Fedorov, A. V., Harper, S. L., Philander, S. G., Winter, B., & Wittenberg, A.(2003). How Predictable is El Nino? Bull. Amer. Meteor. Soc., 84 (7), 911-920. doi: 10.1175/BAMS-84-7-911

Fedorov, A. V., Hu, S., Lengaigne, M., & Guilyardi, E. (2015, Mar 01). The im-pact of westerly wind bursts and ocean initial state on the development,and diversity of El Nino events. Climate Dyn., 44 (5), 1381–1401. doi:10.1007/s00382-014-2126-4

Flugel, M., Chang, P., & Penland, C. (2004). The Role of Stochastic Forcing inModulating ENSO Predictability. J. Climate, 17 (16), 3125-3140. doi: 10.1175/1520-0442(2004)017h3125:TROSFIi2.0.CO;2

Gonzalez, P. L. M., & Goddard, L. (2016). Long-lead ENSO predictability fromCMIP5 decadal hindcasts. Climate Dyn., 46 (9), 3127–3147. doi: 10.1007/s00382-015-2757-0

Gordon, C., Neelin, J. D., Charnock, H., & Philander, S. G. H. (1989). Tropical-ocean-atmosphere interactions in a coupled model. Philos. Trans. Royal Soc.A, 329 (1604), 207-223. doi: 10.1098/rsta.1989.0071

Graham, R. J., Evans, A. D. L., Mylne, K. R., Harrison, M. S. J., & Robertson,K. B. (2000). An assessment of seasonal predictability using atmospheric gen-

–3–

Page 26: El Niño-Southern Oscillation (ENSO) in a Changing Climate ... · 104 the “Climate Diagnostics Bulletin.” Initially, the forecast forum of the Bulletin featured 105 forecasts

manuscript submitted to AGU Books

eral circulation models. Quart. J. Roy. Meteor. Soc., 126 (567), 2211-2240. doi:10.1002/qj.49712656712

Ham, Y.-G., Kug, J.-S., & Park, J.-Y. (2013). Two distinct roles of Atlantic SSTsin ENSO variability: North Tropical Atlantic SST and Atlantic Nino. Geophys.Res. Lett., 40 (15), 4012-4017. doi: 10.1002/grl.50729

Hendon, H. H., Lim, E., Wang, G., Alves, O., & Hudson, D. (2018). Prospects forpredicting two flavors of el nino. Geophys. Res. Lett., 36 (19). doi: 10.1029/2009GL040100

Horsfall, F. (2006). Catalogue of Indices and Definitions of El Nino and La Nina inOperational Use by WMO Members. World Meteorological Organization Com-mission for Climatology.

Hu, S., & Fedorov, A. V. (2016). Exceptionally strong easterly wind burst stalling ElNino of 2014. Proceedings of the National Academy of Sciences, 113 (8), 2005–2010. doi: 10.1073/pnas.1514182113

Huang, B., Shin, C.-S., Shukla, J., Marx, L., Balmaseda, M. A., Halder, S., . . .Kinter, J. L. (2017, 2018/11/14). Reforecasting the ENSO Events inthe Past 57 Years (1958–2014). J. Climate, 30 (19), 7669–7693. doi:10.1175/JCLI-D-16-0642.1

Hudson, D. (2017). ACCESS-S1: The new Bureau of Meteorology multi-week to sea-sonal prediction system. J So Hemisph Earth, 67 (3), 132-159.

Imada, Y., Tatebe, H., Ishii, M., Chikamoto, Y., Mori, M., Arai, M., . . . Kimoto, M.(2015). Predictability of Two Types of El Nio Assessed Using an ExtendedSeasonal Prediction System by MIROC. Mon. Wea. Rev., 143 (11), 4597-4617.doi: 10.1175/MWR-D-15-0007.1

Ineson, S., Balmaseda, M. A., Davey, M. K., Decremer, D., Dunstone, N. J., Gordon,M., . . . Weisheimer, A. (2018). Predicting El Nino in 2014 and 2015. ScientificReports, 8 (1), 10733. doi: 10.1038/s41598-018-29130-1

Inoue, M., & O’Brien, J. J. (1984). A Forecasting Model for the Onset of a MajorEl Nino. Mon. Wea. Rev., 112 (11), 2326-2337. doi: 10.1175/1520-0493(1984)112h2326:AFMFTOi2.0.CO;2

International Research Institute for Climate and Society. (2002). IRI/CPCPlume Forecasts. Retrieved 2019-03-25, from https://iri.columbia.edu/

~forecast/ensofcst/Data/ensoforecast data.txtIzumo, T., Vialard, J., Lengaigne, M., de Boyer Montegut, C., Behera, S. K., Luo,

J.-J., . . . Yamagata, T. (2010, 02). Influence of the state of the Indian OceanDipole on the following year’s El Nino. Nature Geosci , 3 , 168 EP -. Retrievedfrom https://doi.org/10.1038/ngeo760

Jin, E. K., Kinter, J. L., Wang, B., Park, C.-K., Kang, I.-S., Kirtman, B. P., . . .Yamagata, T. (2008, Nov 01). Current status of ENSO prediction skillin coupled ocean–atmosphere models. Climate Dyn., 31 (6), 647–664. doi:10.1007/s00382-008-0397-3

Jin, F.-F. (1997, 2018/11/14). An Equatorial Ocean Recharge Paradigm for ENSO.Part I: Conceptual Model. J. Atmos. Sci., 54 (7), 811–829. doi: 10.1175/1520-0469(1997)054h0811:AEORPFi2.0.CO;2

Karspeck, A. R., Kaplan, A., & Cane, M. A. (2006, 2018/11/14). Predictability Lossin an Intermediate ENSO Model due to Initial Error and Atmospheric Noise.J. Climate, 19 (15), 3572–3588. Retrieved from https://doi.org/10.1175/JCLI3818.1 doi: 10.1175/JCLI3818.1

Katz, R. W. (2002, 05). Sir Gilbert Walker and a Connection between El NinoandStatistics. Stat. Sci., 17 (1), 97–112. doi: 10.1214/ss/1023799000

Keenlyside, N. S., Ding, H., & Latif, M. (2013). Potential of equatorial Atlantic vari-ability to enhance El Nino prediction. Geophys. Res. Lett., 40 (10), 2278-2283.doi: 10.1002/grl.50362

Kirtman, B. P., Min, D., Infanti, J. M., James L. Kinter, I., Paolino, D. A., Zhang,Q., . . . Wood, E. F. (2014). The North American Multimodel Ensemble:

–4–

Page 27: El Niño-Southern Oscillation (ENSO) in a Changing Climate ... · 104 the “Climate Diagnostics Bulletin.” Initially, the forecast forum of the Bulletin featured 105 forecasts

manuscript submitted to AGU Books

Phase-1 Seasonal-to-Interannual Prediction; Phase-2 toward DevelopingIntraseasonal Prediction. Bull. Amer. Meteor. Soc., 95 (4), 585-601. doi:10.1175/BAMS-D-12-00050.1

Kirtman, B. P., Shukla, J., Balmaseda, M., Graham, N., Penland, C., Xue, Y., &Zebiak, S. (2001). Current Status of ENSO Forecast Skill. InternationalCLIVAR Publication Series(56), 1-26.

Kumar, A., Chen, M., Xue, Y., & Behringer, D. (2015, 2018/11/14). An Analysis ofthe Temporal Evolution of ENSO Prediction Skill in the Context of the Equa-torial Pacific Ocean Observing System. Mon. Wea. Rev., 143 (8), 3204–3213.doi: 10.1175/MWR-D-15-0035.1

Kumar, A., Hoerling, M., Ji, M., Leetmaa, A., & Sardeshmukh, P. (1996). Assessinga GCM’s Suitability for Making Seasonal Predictions. J. Climate, 9 (1), 115-129. doi: 10.1175/1520-0442(1996)009h0115:AAGSFMi2.0.CO;2

Larson, S. M., & Kirtman, B. P. (2014). The Pacific Meridional Mode as an ENSOPrecursor and Predictor in the North American Multimodel Ensemble. J. Cli-mate, 27 (18), 7018-7032. doi: 10.1175/JCLI-D-14-00055.1

Larson, S. M., & Kirtman, B. P. (2015). An alternate approach to ensemble ENSOforecast spread: Application to the 2014 forecast. Geophys. Res. Lett., 42 (21),9411-9415. doi: 10.1002/2015GL066173

Larson, S. M., Pegion, K. V., & Kirtman, B. P. (2018). The South Pacific Merid-ional Mode as a Thermally Driven Source of ENSO Amplitude Modulation andUncertainty. J. Climate, 31 (13), 5127-5145. doi: 10.1175/JCLI-D-17-0722.1

Latif, M., Anderson, D., Barnett, T., Cane, M., Kleeman, R., Leetmaa, A., . . .Schneider, E. (1998). A review of the predictability and prediction of ENSO.J. Geophys. Res., 103 (C7), 14375-14393. doi: 10.1029/97JC03413

Lee, J.-W., Yeh, S.-W., & Jo, H.-S. (2016). Weather noise leading to El Nino diver-sity in an ocean general circulation model. Climate Dyn.. doi: 10.1007/s00382-016-3438-3

Lee, T., & McPhaden, M. J. (2010, 2018/11/14). Increasing intensity of El Ninoin the central-equatorial Pacific. Geophys. Res. Lett., 37 (14). doi: 10.1029/2010GL044007

Lengaigne, M., Guilyardi, E., Boulanger, J.-P., Menkes, C., Delecluse, P., Inness,P., . . . Slingo, J. (2004). Triggering of El Nino by westerly wind events ina coupled general circulation model. Climate Dyn., 23 (6), 601–620. doi:10.1007/s00382-004-0457-2

Levine, A., Jin, F. F., & McPhaden, M. J. (2016). Extreme Noise-Extreme El Nino:How State-Dependent Noise Forcing Creates El Nino-La Nina Asymmetry. J.Climate, 29 (15), 5483-5499. doi: 10.1175/JCLI-D-16-0091.1

Levine, A. F. Z., & Jin, F.-F. (2010, 2018/11/14). Noise-Induced Instability in theENSO Recharge Oscillator. J. Atmos. Sci., 67 (2), 529–542. doi: 10.1175/2009JAS3213.1

Levine, A. F. Z., & Jin, F. F. (2017, Jan 01). A simple approach to quantifyingthe noise-ENSO interaction. Part I: deducing the state-dependency of thewindstress forcing using monthly mean data. Climate Dyn., 48 (1), 1–18. doi:10.1007/s00382-015-2748-1

Levine, A. F. Z., & McPhaden, M. J. (2015, 2018/11/14). The annual cycle inENSO growth rate as a cause of the spring predictability barrier. Geophys.Res. Lett., 42 (12), 5034–5041. doi: 10.1002/2015GL064309

Levine, A. F. Z., & McPhaden, M. J. (2016). How the July 2014 easterly windburst gave the 2015-2016 El Nino a head start. Geophys. Res. Lett., 43 (12),6503-6510. doi: 10.1002/2016GL069204

L’Heureux, M. (2015). An overview of the El Nino-Southern Oscillation (ENSO)since 2014. National Weather Service Climate Prediction S&T Digest. doi: 10.7289/V5RN35VW

L’Heureux, M. L., Takahashi, K., Watkins, A. B., Barnston, A. G., Becker, E. J.,

–5–

Page 28: El Niño-Southern Oscillation (ENSO) in a Changing Climate ... · 104 the “Climate Diagnostics Bulletin.” Initially, the forecast forum of the Bulletin featured 105 forecasts

manuscript submitted to AGU Books

Di Liberto, T. E., . . . Wittenberg, A. T. (2017). Observing and Predict-ing the 2015/16 El Nino. Bull. Amer. Meteor. Soc., 98 (7), 1363-1382. doi:10.1175/BAMS-D-16-0009.1

L’Heureux, M. L., Tippett, M. K., & Barnston, A. G. (2015). Character-izing ENSO coupled variability and its impact on North American sea-sonal precipitation and temperature. J. Climate, 28 (10), 4231-4245. doi:10.1175/JCLI-D-14-00508.1

L’Heureux, M. L., Tippett, M. K., Takahashi, K., Barnston, A. G., Becker,E. J., Bell, G. D., . . . Wang, W. (2019). Strength Outlooks for theEl Nino-Southern Oscillation. Wea. Forecasting , 34 (1), 165-175. doi:10.1175/WAF-D-18-0126.1

Liu, Z., Jin, Y., & Rong, X. (2019). A Theory for the Seasonal Predictability Bar-rier: Threshold, Timing, and Intensity. J. Climate, 32 (2), 423-443. doi: 10.1175/JCLI-D-18-0383.1

Lopez, H., & Kirtman, B. P. (2014, 2018/11/14). WWBs, ENSO predictability,the spring barrier and extreme events. J. Geophys. Res., 119 (17), 10,114–10,138. Retrieved from https://doi.org/10.1002/2014JD021908 doi:10.1002/2014JD021908

Luo, J.-J., Liu, G., Hendon, H., Alves, O., & Yamagata, T. (2017). Inter-basinsources for two-year predictability of the multi-year La Nina event in 2010–2012. Sci. Rep., 7 (1), 2276. doi: 10.1038/s41598-017-01479-9

Luo, J.-J., Masson, S., Behera, S. K., & Yamagata, T. (2008). Extended ENSO Pre-dictions Using a Fully Coupled Ocean-Atmosphere Model. J. Climate, 21 (1),84-93. doi: 10.1175/2007JCLI1412.1

Luo, J.-J., Yuan, C., Sasaki, W., Behera, S. K., Masumoto, Y., Yamagata, T.,. . . Masson, S. (2016). CURRENT STATUS OF INTRASEASONAL-SEASONAL-TO-INTERANNUAL PREDICTION OF THE INDO-PACIFICCLIMATE. In Indo-Pacific Climate Variability and Predictability (p. 63-107).doi: 10.1142/9789814696623 0003

Luo, J.-J., Zhang, R., Behera, S. K., Masumoto, Y., Jin, F.-F., Lukas, R., & Yama-gata, T. (2010, 2018/11/14). Interaction between El Nino and Extreme IndianOcean Dipole. J. Climate, 23 (3), 726–742. doi: 10.1175/2009JCLI3104.1

MacLachlan, C., Arribas, A., Peterson, K. A., Maidens, A., Fereday, D., Scaife,A. A., . . . Madec, G. (2015). Global Seasonal forecast system version 5(GloSea5): a high-resolution seasonal forecast system. Quart. J. Roy. Meteor.Soc., 141 (689), 1072-1084. doi: 10.1002/qj.2396

Magnusson, L., Alonso-Balmaseda, M., Corti, S., Molteni, F., & Stockdale, T. (2013,Nov 01). Evaluation of forecast strategies for seasonal and decadal forecasts inpresence of systematic model errors. Climate Dyn., 41 (9), 2393–2409. doi: 10.1007/s00382-012-1599-2

Martin, G. M., Milton, S. F., Senior, C. A., Brooks, M. E., Ineson, S., Reichler, T.,& Kim, J. (2010). Analysis and Reduction of Systematic Errors through aSeamless Approach to Modeling Weather and Climate. J. Climate, 23 (22),5933-5957. doi: 10.1175/2010JCLI3541.1

Martın-Rey, M., Rodrıguez-Fonseca, B., Polo, I., & Kucharski, F. (2014). On theAtlantic–Pacific Ninos connection: a multidecadal modulated mode. ClimateDyn., 43 (11), 3163–3178. doi: 10.1007/s00382-014-2305-3

McPhaden, M. J. (2003, 2018/11/14). Tropical Pacific Ocean heat content varia-tions and ENSO persistence barriers. Geophys. Res. Lett., 30 (9). doi: 10.1029/2003GL016872

McPhaden, M. J. (2012, 2018/11/14). A 21st century shift in the relationship be-tween enso sst and warm water volume anomalies. Geophys. Res. Lett., 39 (9).doi: 10.1029/2012GL051826

McPhaden, M. J. (2015, 09). Playing hide and seek with El Nino. Nature Clim.Change, 5 (9), 791–795. doi: 10.1038/nclimate2775

–6–

Page 29: El Niño-Southern Oscillation (ENSO) in a Changing Climate ... · 104 the “Climate Diagnostics Bulletin.” Initially, the forecast forum of the Bulletin featured 105 forecasts

manuscript submitted to AGU Books

McPhaden, M. J., Busalacchi, A. J., & Anderson, D. L. (2010, September). ATOGA Retrospective. Oceanography , 23 . Retrieved from https://doi.org/10.5670/oceanog.2010.26

McPhaden, M. J., Busalacchi, A. J., Cheney, R., Donguy, J.-R., Gage, K. S.,Halpern, D., . . . Takeuchi, K. (1998). The Tropical Ocean-Global Atmo-sphere observing system: A decade of progress. J. Geophys. Res., 103 (C7),14169-14240. doi: 10.1029/97JC02906

McPhaden, M. J., Timmermann, A., Widlansky, M. J., Balmaseda, M. A., &Stockdale, T. N. (2015). The Curious Case of the El Nino That NeverHappened: A Perspective from 40 Years of Progress in Climate Researchand Forecasting. Bull. Amer. Meteor. Soc., 96 (10), 1647-1665. doi:10.1175/BAMS-D-14-00089.1

Meinen, C. S., & McPhaden, M. J. (2000, 2018/11/14). Observations ofWarm Water Volume Changes in the Equatorial Pacific and Their Rela-tionship to El Nino and La Nina. J. Climate, 13 (20), 3551–3559. doi:10.1175/1520-0442(2000)013h3551:OOWWVCi2.0.CO;2

Menkes, C. E., Lengaigne, M., Vialard, J., Puy, M., Marchesiello, P., Cravatte, S., &Cambon, G. (2014). About the role of Westerly Wind Events in the possibledevelopment of an El Nino in 2014. Geophys. Res. Lett., 41 (18), 6476-6483.doi: 10.1002/2014GL061186

Min, Q., Su, J., Zhang, R., & Rong, X. (2015). What hindered the El Nio pattern in2014? Geophys. Res. Lett., 42 (16), 6762-6770. doi: 10.1002/2015GL064899

Min, Y.-M., Kryjov, V. N., & Park, C.-K. (2009). A Probabilistic Multimodel En-semble Approach to Seasonal Prediction. Wea. Forecasting , 24 (3), 812-828.doi: 10.1175/2008WAF2222140.1

National Weather Service. (2016). Technical Implementation Notice 16-09. Re-trieved 2019-03-25, from http://www.nws.noaa.gov/os/notification/tin16-09cfs.htm

Neske, S., & McGregor, S. (2018, 2018/11/14). Understanding the Warm Water Vol-ume Precursor of ENSO Events and its Interdecadal Variation. Geophys. Res.Lett., 45 (3), 1577–1585. doi: 10.1002/2017GL076439

Newman, M. (2013, 2018/11/14). An Empirical Benchmark for Decadal Forecasts ofGlobal Surface Temperature Anomalies. J. Climate, 26 (14), 5260–5269. doi:10.1175/JCLI-D-12-00590.1

Newman, M., & Sardeshmukh, P. D. (2017). Are we near the predictability limitof tropical Indo-Pacific sea surface temperatures? Geophys. Res. Lett., 44 (16),8520-8529. doi: 10.1002/2017GL074088

Newman, M., Shin, S.-I., & Alexander, M. A. (2011, 2018/11/14). Natural variationin ENSO flavors. Geophys. Res. Lett., 38 (14). doi: 10.1029/2011GL047658

Newman, M., Wittenberg, A. T., Cheng, L., Compo, G. P., & Smith, C. A. (2018).The Extreme 2015/16 El Nino, in the Context of Historical Climate Vari-ability and Change. Bull. Amer. Meteor. Soc., 99 (1), S16-S20. doi:10.1175/BAMS-D-17-0116.1

Orlove, B. S., Chiang, J. C. H., & Cane, M. A. (2000). Forecasting Andean rain-fall and crop yield from the influence of El Nino on Pleiades visibility. Nature,403 (6765), 68–71. doi: 10.1038/47456

Palmer, T. N., Doblas-Reyes, F. J., Hagedorn, R., Alessandri, A., Gualdi, S., Ander-sen, U., . . . Thomson, M. C. (2004). Development of a European multimodelensemble system for seasonal-to-interannual prediction (DEMETER). Bull.Amer. Meteor. Soc., 85 (6), 853-872. doi: 10.1175/BAMS-85-6-853

Pegion, K. V., & Selman, C. (2017). Extratropical precursors of the el nino-southernoscillation. In Climate extremes (p. 299-314). American Geophysical Union(AGU). doi: 10.1002/9781119068020.ch18

Planton, Y., Vialard, J., Guilyardi, E., Lengaigne, M., & Izumo, T. (2018). WesternPacific Oceanic Heat Content: A Better Predictor of La Nina Than of El Nino.

–7–

Page 30: El Niño-Southern Oscillation (ENSO) in a Changing Climate ... · 104 the “Climate Diagnostics Bulletin.” Initially, the forecast forum of the Bulletin featured 105 forecasts

manuscript submitted to AGU Books

Geophys. Res. Lett., 45 (18), 9824-9833. doi: 10.1029/2018GL079341Puy, M., Vialard, J., Lengaigne, M., & Guilyardi, E. (2016). Modulation of equa-

torial Pacific westerly/easterly wind events by the Madden–Julian oscillationand convectively-coupled Rossby waves. Climate Dyn., 46 (7), 2155–2178. doi:10.1007/s00382-015-2695-x

Quinn, W. H. (1974a). Monitoring and Predicting El Nino Invasions. J. Appl.Meteor. Climatol., 13 (7), 825-830. doi: 10.1175/1520-0450(1974)013h0825:MAPENIi2.0.CO;2

Quinn, W. H. (1974b). Outlook for El Nino-like conditions in 1975. NORPAX High-lights, 2 (6), 2-3.

Reeves, R. W., & Gemmill, D. J. (2004). Reflections on 25 Years of Analysis, Diag-nosis, and Prediction, 1979-2004. U.S. Government Printing O�ce.

Ren, H.-L., & Jin, F.-F. (2013, 2018/11/14). Recharge Oscillator Mechanisms inTwo Types of ENSO. J. Climate, 26 (17), 6506–6523. doi: 10.1175/JCLI-D-12-00601.1

Roulston, M. S., & Neelin, J. D. (2000, 2018/11/14). The response of an ENSOModel to climate noise, weather noise and intraseasonal forcing. Geophys. Res.Lett., 27 (22), 3723–3726. doi: 10.1029/2000GL011941

Santoso, A., Hendon, H., Watkins, A., Power, S., Dommenget, D., England, M., . . .Delage, F. (2018). Dynamics and predictability of the El Nino-Southern Os-cillation: An Australian perspective on progress and challenges. Bull. Amer.Meteor. Soc., 0 (0), null. doi: 10.1175/BAMS-D-18-0057.1

Santoso, A., Mcphaden, M. J., & Cai, W. (2017). The Defining Characteristicsof ENSO Extremes and the Strong 2015/2016 El Nino. Rev. Geophys., 55 (4),1079-1129. doi: 10.1002/2017RG000560

Stein, K., Schneider, N., Timmermann, A., & Jin, F.-F. (2010, 2018/11/14). Sea-sonal Synchronization of ENSO Events in a Linear Stochastic Model. J. Cli-mate, 23 (21), 5629–5643. doi: 10.1175/2010JCLI3292.1

Stein, K., Timmermann, A., Schneider, N., Jin, F.-F., & Stuecker, M. F. (2014,2018/11/14). ENSO Seasonal Synchronization Theory. J. Climate, 27 (14),5285–5310. doi: 10.1175/JCLI-D-13-00525.1

Stephenson, D., Coelho, C., Doblas-Reyes, F., & Balmaseda, M. (2005). Fore-cast assimilation: a unified framework for the combination of multi-modelweather and climate predictions. TELLUS A, 57 (3), 253-264. doi:10.3402/tellusa.v57i3.14664

Stevenson, S., Capotondi, A., Fasullo, J., & Otto-Bliesner, B. (2017). Forcedchanges to twentieth century ENSO diversity in a last Millennium context.Climate Dyn.. doi: 10.1007/s00382-017-3573-5

Stockdale, T., Doblas-Reyes, F., & Ferranti, L. (2009). EUROSIP: multi-model sea-sonal forecasting [Meteorology]. ECMWF Newsletter , 10-16. Retrieved fromhttps://www.ecmwf.int/node/17491 doi: 10.21957/7wc0nybvir

Stockdale, T. N., Anderson, D. L. T., Alves, J. O. S., & Balmaseda, M. A. (1998).Global seasonal rainfall forecasts using a coupled ocean–atmosphere model.Nature, 392 (6674), 370–373. Retrieved from https://doi.org/10.1038/32861 doi: 10.1038/32861

Takahashi, K., & Dewitte, B. (2016, Mar 01). Strong and moderate nonlinear ElNino regimes. Climate Dyn., 46 (5), 1627–1645. doi: 10.1007/s00382-015-2665-3

Thomas, E. E., Vimont, D. J., Newman, M., Penland, C., & Martınez-Villalobos,C. (2018, 2018/11/14). The Role of Stochastic Forcing in Generating ENSODiversity. J. Climate, 31 (22), 9125–9150. doi: 10.1175/JCLI-D-17-0582.1

Timmermann, A., An, S.-I., Kug, J.-S., Jin, F.-F., Cai, W., Capotondi, A., . . .Zhang, X. (2018). El nino–southern oscillation complexity. Nature, 559 (7715),535–545. doi: 10.1038/s41586-018-0252-6

Tippett, M. K., & Barnston, A. G. (2008). Skill of Multimodel ENSO Probability

–8–

Page 31: El Niño-Southern Oscillation (ENSO) in a Changing Climate ... · 104 the “Climate Diagnostics Bulletin.” Initially, the forecast forum of the Bulletin featured 105 forecasts

manuscript submitted to AGU Books

Forecasts. Mon. Wea. Rev., 136 (10), 3933-3946. doi: 10.1175/2008MWR2431.1

Tippett, M. K., Barnston, A. G., & Li, S. (2012). Performance of Recent Multi-model ENSO Forecasts. J. Appl. Meteor. Climatol., 51 (3), 637-654. doi: 10.1175/JAMC-D-11-093.1

Tippett, M. K., Ranganathan, M., L’Heureux, M., Barnston, A. G., & DelSole,T. (2017). Assessing probabilistic predictions of ENSO phase and intensityfrom the North American Multimodel Ensemble. Climate Dyn., 1–22. doi:10.1007/s00382-017-3721-y

Trenberth, K. E. (1997). The Definition of El Nino. Bull. Amer. Meteor. Soc.,78 (12), 2771-2778. doi: 10.1175/1520-0477(1997)078h2771:TDOENOi2.0.CO;2

Tziperman, E., & Yu, L. (2007). Quantifying the Dependence of Westerly WindBursts on the Large-Scale Tropical Pacific SST. J. Climate, 20 (12), 2760-2768.doi: 10.1175/JCLI4138a.1

van Oldenborgh, G. J., Balmaseda, M. A., Ferranti, L., Stockdale, T. N., & Ander-son, D. L. T. (2005a). Did the ECMWF Seasonal Forecast Model OutperformStatistical ENSO Forecast Models over the Last 15 Years? J. Climate, 18 (16),3240-3249. doi: 10.1175/JCLI3420.1

van Oldenborgh, G. J., Balmaseda, M. A., Ferranti, L., Stockdale, T. N., & Ander-son, D. L. T. (2005b). Evaluation of Atmospheric Fields from the ECMWFSeasonal Forecasts over a 15-Year Period. Journal of Climate, 18 (16), 3250-3269. doi: 10.1175/JCLI3421.1

Vimont, D. J. (2010, 2018/11/14). Transient Growth of Thermodynamically Cou-pled Variations in the Tropics under an Equatorially Symmetric Mean State.J. Climate, 23 (21), 5771–5789. doi: 10.1175/2010JCLI3532.1

Vimont, D. J., Battisti, D. S., & Hirst, A. C. (2001). Footprinting: A seasonalconnection between the tropics and mid-latitudes. Geophys. Res. Lett., 28 (20),3923-3926. doi: 10.1029/2001GL013435

Vimont, D. J., Battisti, D. S., & Hirst, A. C. (2003, 2018/11/14). The Sea-sonal Footprinting Mechanism in the CSIRO General Circulation Models.J. Climate, 16 (16), 2653–2667. doi: 10.1175/1520-0442(2003)016h2653:TSFMITi2.0.CO;2

Vimont, D. J., Wallace, J. M., & Battisti, D. S. (2003, 2018/11/14). The SeasonalFootprinting Mechanism in the Pacific: Implications for ENSO. J. Climate,16 (16), 2668–2675. doi: 10.1175/1520-0442(2003)016h2668:TSFMITi2.0.CO;2

Wang, G., & Hendon, H. H. (2017). Why 2015 was a strong El Nino and 2014 wasnot. Geophys. Res. Lett., 44 (16), 8567-8575. doi: 10.1002/2017GL074244

Wang, S.-Y., L’Heureux, M., & Chia, H.-H. (2012). ENSO prediction one year inadvance using western North Pacific sea surface temperatures. Geophys. Res.Lett., 39 (5). doi: 10.1029/2012GL050909

Weisheimer, A., Doblas-Reyes, F. J., Palmer, T. N., Alessandri, A., Arribas, A.,Dqu, M., . . . Rogel, P. (2009). ENSEMBLES: A new multi-model ensem-ble for seasonal-to-annual predictions- Skill and progress beyond DEME-TER in forecasting tropical Pacific SSTs. Geophys. Res. Lett., 36 (21). doi:10.1029/2009GL040896

Wittenberg, A. T., Rosati, A., Delworth, T. L., Vecchi, G. A., & Zeng, F. (2014,2018/11/14). ENSO Modulation: Is It Decadally Predictable? J. Climate,27 (7), 2667–2681. doi: 10.1175/JCLI-D-13-00577.1

Xue, Y., Chen, M., Kumar, A., Hu, Z.-Z., & Wang, W. (2013). Prediction Skilland Bias of Tropical Pacific Sea Surface Temperatures in the NCEP Cli-mate Forecast System Version 2. J. Climate, 26 (15), 5358-5378. doi:10.1175/JCLI-D-12-00600.1

Yeh, S.-W., Kug, J.-S., Dewitte, B., Kwon, M.-H., Kirtman, B. P., & Jin, F.-F.(2009, 09 24). El Nino in a changing climate. Nature, 461 , 511 EP -.

–9–

Page 32: El Niño-Southern Oscillation (ENSO) in a Changing Climate ... · 104 the “Climate Diagnostics Bulletin.” Initially, the forecast forum of the Bulletin featured 105 forecasts

manuscript submitted to AGU Books

You, Y., & Furtado, J. C. (2017). The role of South Pacific atmospheric variabil-ity in the development of di↵erent types of ENSO. Geophys. Res. Lett., 44 (14),7438-7446. doi: 10.1002/2017GL073475

Yu, Y., Mu, M., & Duan, W. (2012). Does model parameter error cause a significant?spring predictability barrier? for el nio events in the zebiak?cane model? J.Climate, 25 (4), 1263-1277. doi: 10.1175/2011JCLI4022.1

Zebiak, S. E., & Cane, M. A. (1987). A Model El Nino?Southern Oscillation.Mon. Wea. Rev., 115 (10), 2262-2278. doi: 10.1175/1520-0493(1987)115h2262:AMENOi2.0.CO;2

Zhang, H., Clement, A., & Di Nezio, P. (2014). The South Pacific Meridional Mode:A Mechanism for ENSO-like Variability. J. Climate, 27 (2), 769-783. doi: 10.1175/JCLI-D-13-00082.1

Zhao, M., Hendon, H. H., Alves, O., Liu, G., & Wang, G. (2016, 2018/11/14).Weakened Eastern Pacific El Nino Predictability in the Early Twenty-FirstCentury. J. Climate, 29 (18), 6805–6822. doi: 10.1175/JCLI-D-15-0876.1

Zheng, F., & Yu, J.-Y. (2017, Dec 01). Contrasting the skills and biases of deter-ministic predictions for the two types of El Nino. Advances in AtmosphericSciences, 34 (12), 1395–1403. doi: 10.1007/s00376-017-6324-y

Zheng, F., & Zhu, J. (2010). Coupled assimilation for an intermediated coupledENSO prediction model. Ocean Dynam., 60 (5), 1061–1073. doi: 10.1007/s10236-010-0307-1

Zhu, J., Kumar, A., Huang, B., Balmaseda, M. A., Hu, Z.-Z., Marx, L., & Kinter III,J. L. (2016, 01 20). The role of o↵-equatorial surface temperature anomalies inthe 2014 El Nino prediction. Sci. Rep., 6 , 19677 EP -.

–10–

Page 33: El Niño-Southern Oscillation (ENSO) in a Changing Climate ... · 104 the “Climate Diagnostics Bulletin.” Initially, the forecast forum of the Bulletin featured 105 forecasts

For Review Only

Figure 1. A timeline of events in ENSO forecasting.

Page 33 of 44 AGU Books

Page 34: El Niño-Southern Oscillation (ENSO) in a Changing Climate ... · 104 the “Climate Diagnostics Bulletin.” Initially, the forecast forum of the Bulletin featured 105 forecasts

For Review Only

Figure 2. The first ENSO Diagnostics Advisory issued at NOAA Climate Prediction Center in March 1986, which assessed an emerging warm episode. Photo Credit: Robert Reeves.

883x1319mm (72 x 72 DPI)

Page 34 of 44AGU Books

Page 35: El Niño-Southern Oscillation (ENSO) in a Changing Climate ... · 104 the “Climate Diagnostics Bulletin.” Initially, the forecast forum of the Bulletin featured 105 forecasts

For Review Only

Figure 3. The forecast lead-time (y-axis) when Niño-3.4 index skill exceeds certain correlation thresholds. On the x-axis, the target month of the forecast is presented. Model data is based on the ensemble average

of ~100 members from the North American Multi-Model Ensemble from 1982 to 2018. Departures in the Niño-3.4 index are based on 1982-2010 monthly averages.

Page 35 of 44 AGU Books

Page 36: El Niño-Southern Oscillation (ENSO) in a Changing Climate ... · 104 the “Climate Diagnostics Bulletin.” Initially, the forecast forum of the Bulletin featured 105 forecasts

For Review Only

Figure 4. Predictions of the Niño-3.4 index from the North American Multi-Model Ensemble initialized at the beginning of February 2019. Predictions are shown out to 12 months lead-time or through January 2020.

The black solid line are the observations based on daily OISST, and the dashed line is the ensemble average. Remaining lines show the individual members. Departures in the Niño-3.4 index are based on

1982-2010 monthly averages.

Page 36 of 44AGU Books

Page 37: El Niño-Southern Oscillation (ENSO) in a Changing Climate ... · 104 the “Climate Diagnostics Bulletin.” Initially, the forecast forum of the Bulletin featured 105 forecasts

For Review Only

Figure 5. Year-to-year variations of model hindcast/forecast skill in the tropical Pacific ENSO region for monthly forecasts initialized during 1961-2015, for six month forecast lead time. The measure of skill used

is the pattern correlation between ensemble mean forecast and corresponding observed monthly SST anomalies within the region bounded by 20ºS-20ºN, 170ºE-70ºW, smoothed with a 13-month running mean

filter to emphasize annual variations. All anomalies are determined relative to a 1961-2015 monthly climatology. The light red (blue) shaded epochs indicate periods of >1 sigma warm (< −1 sigma cold) Niño3.4 seasonal SST anomalies. Skill is shown for the Linear Inverse Model (LIM), the North American

Multi-Model Ensemble (NMME) operational ensemble, the ECMWF SEAS5 ensemble, the JAMSTEC SINTEX-F ensemble, and model-analogs employed on either a 4-model subset of the NMME models or a 10-model

subset of the CMIP5 models. The mean expected skill (a predictability estimate) from the LIM is also shown. The LIM is trained from tropical sea surface analyses over the 1958-2011 period, with hindcasts determined by a ten-fold cross validation technique. The NMME and SEAS5 hindcasts are bias corrected over the 1982-

2009 period. Therefore, all forecasts after 2011 for the LIM, NMME, SEAS5, and SINTEX-F models are independent of observations during that period. There is no training period for the model-analog forecasts, since they are based on analogs to tropical IndoPacific sea surface observations drawn from independent

multi-centennial unitialized model simulations.

Page 37 of 44 AGU Books

Page 38: El Niño-Southern Oscillation (ENSO) in a Changing Climate ... · 104 the “Climate Diagnostics Bulletin.” Initially, the forecast forum of the Bulletin featured 105 forecasts

For Review Only

Figure 6. (top panel) Predictions of the Niño-3.4 index from the North American Multi-Model Ensemble initialized at the beginning of April 2014. Predictions are shown out to 12 months lead-time or through March 2015. The black solid line are the observations based on daily OISST, and the dashed line is the

ensemble average. Remaining lines show the individual members. Departures in the Niño-3.4 index are based on 1982-2010 monthly averages. (bottom panel) Predictions of the Niño-3.4 index from the EUROSIP multi-model system initialized on the 1st April 2019, according to the operational calibrated product. Various

percentiles of the distribution are shown, based on a multi-model combination of variance-corrected ensemble forecasts, with an additional calibration of the forecast uncertainty to match historic forecast error.

Page 38 of 44AGU Books

Page 39: El Niño-Southern Oscillation (ENSO) in a Changing Climate ... · 104 the “Climate Diagnostics Bulletin.” Initially, the forecast forum of the Bulletin featured 105 forecasts

For Review Only

Figure 7. Log skill score (LSS) from hindcasts (1981-2010) of the NCEP Climate Forecast System (CFS) version 2. Skill is presented as a function of target month (x-axis) and lead-time (y-axis). The top left panel shows the skill of the Niño-3.4 SST index region (5°S-5°N, 170°W-120°W), the bottom left panel

shows the skill of the Equatorial Southern Oscillation Index (surface pressure differences based 5°S-5°N, 80°W-130°W minus 5°S-5°N, 90°E-140°E), the top right panel shows the skill of the Eastern Pacific

Outgoing Longwave Radiation (OLR) index (5°S-5°N, 160°W-110°W; Chiodi and Harrison, 2013), and the bottom right panel shows the skill of the Central Pacific OLR index (5°S-5°N, 170°E-140°W; L’Heureux et

al., 2015). Positive values of LSS show forecasts that are better than a climatological forecast and negative values of LSS show where forecasts are worse than climatology.

Page 39 of 44 AGU Books

Page 40: El Niño-Southern Oscillation (ENSO) in a Changing Climate ... · 104 the “Climate Diagnostics Bulletin.” Initially, the forecast forum of the Bulletin featured 105 forecasts

For Review Only

Figure 1. A timeline of events in ENSO forecasting.

Page 40 of 44AGU Books

Page 41: El Niño-Southern Oscillation (ENSO) in a Changing Climate ... · 104 the “Climate Diagnostics Bulletin.” Initially, the forecast forum of the Bulletin featured 105 forecasts

For Review Only

Figure 3. The forecast lead-time (y-axis) when Niño-3.4 index skill exceeds certain correlation thresholds. On the x-axis, the target month of the forecast is presented. Model data is based on the ensemble average

of ~100 members from the North American Multi-Model Ensemble from 1982 to 2018. Departures in the Niño-3.4 index are based on 1982-2010 monthly averages.

Page 41 of 44 AGU Books

Page 42: El Niño-Southern Oscillation (ENSO) in a Changing Climate ... · 104 the “Climate Diagnostics Bulletin.” Initially, the forecast forum of the Bulletin featured 105 forecasts

For Review Only

Figure 4. Predictions of the Niño-3.4 index from the North American Multi-Model Ensemble initialized at the beginning of February 2019. Predictions are shown out to 12 months lead-time or through January 2020.

The black solid line are the observations based on daily OISST, and the dashed line is the ensemble average. Remaining lines show the individual members. Departures in the Niño-3.4 index are based on

1982-2010 monthly averages.

Page 42 of 44AGU Books

Page 43: El Niño-Southern Oscillation (ENSO) in a Changing Climate ... · 104 the “Climate Diagnostics Bulletin.” Initially, the forecast forum of the Bulletin featured 105 forecasts

For Review Only

Figure 5. Year-to-year variations of model hindcast/forecast skill in the tropical Pacific ENSO region for monthly forecasts initialized during 1961-2015, for six month forecast lead time. The measure of skill used

is the pattern correlation between ensemble mean forecast and corresponding observed monthly SST anomalies within the region bounded by 20ºS-20ºN, 170ºE-70ºW, smoothed with a 13-month running mean

filter to emphasize annual variations. All anomalies are determined relative to a 1961-2015 monthly climatology. The light red (blue) shaded epochs indicate periods of >1 sigma warm (< −1 sigma cold) Niño3.4 seasonal SST anomalies. Skill is shown for the Linear Inverse Model (LIM), the North American

Multi-Model Ensemble (NMME) operational ensemble, the ECMWF SEAS5 ensemble, the JAMSTEC SINTEX-F ensemble, and model-analogs employed on either a 4-model subset of the NMME models or a 10-model

subset of the CMIP5 models. The mean expected skill (a predictability estimate) from the LIM is also shown. The LIM is trained from tropical sea surface analyses over the 1958-2011 period, with hindcasts determined by a ten-fold cross validation technique. The NMME and SEAS5 hindcasts are bias corrected over the 1982-

2009 period. Therefore, all forecasts after 2011 for the LIM, NMME, SEAS5, and SINTEX-F models are independent of observations during that period. There is no training period for the model-analog forecasts, since they are based on analogs to tropical IndoPacific sea surface observations drawn from independent

multi-centennial unitialized model simulations.

Page 43 of 44 AGU Books

Page 44: El Niño-Southern Oscillation (ENSO) in a Changing Climate ... · 104 the “Climate Diagnostics Bulletin.” Initially, the forecast forum of the Bulletin featured 105 forecasts

For Review Only

Figure 6. (top panel) Predictions of the Niño-3.4 index from the North American Multi-Model Ensemble initialized at the beginning of April 2014. Predictions are shown out to 12 months lead-time or through March 2015. The black solid line are the observations based on daily OISST, and the dashed line is the

ensemble average. Remaining lines show the individual members. Departures in the Niño-3.4 index are based on 1982-2010 monthly averages. (bottom panel) Predictions of the Niño-3.4 index from the EUROSIP multi-model system initialized on the 1st April 2019, according to the operational calibrated product. Various

percentiles of the distribution are shown, based on a multi-model combination of variance-corrected ensemble forecasts, with an additional calibration of the forecast uncertainty to match historic forecast error.

Page 44 of 44AGU Books

Page 45: El Niño-Southern Oscillation (ENSO) in a Changing Climate ... · 104 the “Climate Diagnostics Bulletin.” Initially, the forecast forum of the Bulletin featured 105 forecasts

For Review Only

Figure 7. Log skill score (LSS) from hindcasts (1981-2010) of the NCEP Climate Forecast System (CFS) version 2. Skill is presented as a function of target month (x-axis) and lead-time (y-axis). The top left panel shows the skill of the Niño-3.4 SST index region (5°S-5°N, 170°W-120°W), the bottom left panel

shows the skill of the Equatorial Southern Oscillation Index (surface pressure differences based 5°S-5°N, 80°W-130°W minus 5°S-5°N, 90°E-140°E), the top right panel shows the skill of the Eastern Pacific

Outgoing Longwave Radiation (OLR) index (5°S-5°N, 160°W-110°W; Chiodi and Harrison, 2013), and the bottom right panel shows the skill of the Central Pacific OLR index (5°S-5°N, 170°E-140°W; L’Heureux et

al., 2015). Positive values of LSS show forecasts that are better than a climatological forecast and negative values of LSS show where forecasts are worse than climatology.

Page 45 of 44 AGU Books