A systematic comparison of particle filter and EnKF in ...

0 downloads 0 Views 5MB Size Report
The pParticle filter (PF) and the ensemble Kalman filter (EnKF) are two promising ... 2012) or inferred from proxy-based reconstructions (Mann et al., 2008, 2009;.
A systematic comparison of particle filter and EnKF in assimilating time-averaged observations

1 2

3 Huaran Liu1,2, Zhengyu Liu2, and Feiyu Lu2

4

5

1

6

China

7

2

College of Atmospheric Sciences, Nanjing University of Information Science & Technology,

Department of Atmospheric and Oceanic Sciences, University of Wisconsin-Madison, USA.

8 9 10 11 12

Corresponding author: Huaran Liu ([email protected])

13

Key Points:

14

● Pseudo proxy experiments using particle filter with simple importance resampling and

15

ensemble Kalman filter to evaluate the performance for paleoclimate reconstruction

16

● The strengths and weaknesses of these two filters for paleoclimate data assimilation

17

● In sparse observation network, large localization radius is crucial to EnKF

18

19

Abstract

20 21

The pParticle filter (PF) and the ensemble Kalman filter (EnKF) are two promising and popularly

22

adopted types of ensemble-based data assimilation methods for paleoclimate reconstruction.

23

However, no systematic comparison between them has been attempted. We compare these two

24

uncertainty based methods in pseudo proxy experiments where synthetic seasonal mean sea

25

surface temperature (SST) observations are assimilated. Their skills were evaluated with regards

26

to local performance, large scale performance, and their ability to capture the modes of variability.

27

It is found that the EAKF (Ensemble Adjstment Kalman Filter, a variant of EnKF) performs better

28

than the particle filterPF with only one third of the ensemble size, despite particle filter’sPF’s

29

theoretical superiority in allowing for non-Gaussian statistics and nonlinear dynamics. The success

30

of the EAKF is attributed to the existence of large-scale co-variability and hence the adoption of a

31

large localization scale in the experiment, which allows broader areas to be impacted by a sparse

32

observation networkthe fact that: 1) Gaussian assumption is appropriate for this application; 2)The

33

EAKF is less sensitive to sampling errors than the PF due to the different methodological natures:

34

16 members are enough to capture accurate covariance for the EAKF, but 48 (even 96) members

35

still underrepresent the state space of such high-dimensional system for the PF. Our study

36

highlights the importance of using a large localization radius to constrain large-scale patterns for

37

the application of the EnKF to paleoclimate reconstruction given such sparse proxy network,

38

andThis study suggests that additional techniques, such as localization or clustered particle filter,

39

are needed to improve the particle filter for paleoclimate reconstruction, in addition to the simple

40

importance resampling which is currently adopted by most research.

41

42

1 Introduction

43

Past climate can be simulated in the state-of-the-art climate models (Otto-Blieser et al.2006;

44

Braconnot et al. 2012) or inferred from proxy-based reconstructions (Mann et al., 2008, 2009;

45

Marcott et al. 2013). Both approaches provide independent yet complementary information about

46

past climate and have been used independently for the reconstruction and understanding of the

47

past climate. Paleoclimate data assimilation (PDA) offers a new approach to further improve the

48

reconstruction of past climate by combining the observational information from the proxy records

49

and the dynamic constraints from climate models. PDA will constrain the model with proxy

50

observations and provide a complete field of the past climate state containing information that

51

cannot be directly derived from the proxy records, equivalent to reanalysis datasets that fill in the

52

spatial and temporal gaps in a sparse observation network (Kalnay, 2003). Furthermore, with

53

careful calibration of the uncertainty in the model and proxy data, PDA can also quantify the

54

uncertainty of the final analysis product.

55

In contrast to data assimilation (DA) techniques in meteorology, PDA is still in its infancy

56

and no universal methodology has been established. The difficulties posed by spatially and

57

temporally sparse, noisy and mostly indirect proxy records make it challenging to apply most

58

modern DA techniques directly to PDA. Various DA methods have been attempted since the

59

concept was introduced to the paleo field by Von Storch (2000). Early methods include pattern

60

nudging (von Storch et al., 2000) and forcing singular vectors (Barkmeijer et al., 2003). These

61

methods share the same idea by adding an artificial forcing term to the prognostic equations to

62

nudge the model towards a desired large-scale pattern derived from available observations. The

63

drawbacks of these methods are that they require a relatively dense observation network to capture

64

the large-scale pattern and that they cannot address the uncertainty of the analysis. Recently,

65

ensemble filter methods have been widely used for paleoclimate studies, in which a finite ensemble

66

of model simulations are used to represent the model’s statistical behavior and Bayesian estimation

67

theory is used to derive the posterior estimates (Leeuwen 2009; Evensen, 1994, 2003). These

68

ensemble methods fall into two categories: the particle filter and the ensemble-based Kalman filter.

69

The application of the particle filter to paleoclimate has experienced two three stages. In

70

the first stage, a simplified (off-line) particle filter was used and a single member that best matches

71

the proxy record is selected among an ensemble of pre-existing model simulations as the final

72

analysis (Goosse et al., 2006a, 2006b), hence no forwarding model is needed. In the second stage,

73

this method was modified towards an on-line degenerate particle filter; (Goosse et al. 2009, 2010a;

74

Crespin et al., 2009; Widmann et al., 2010), in which one member was selected as the best-fit of

75

the observations at the target period and is then used as the initial conditions for the subsequent

76

simulation. In the second third stage, the on-line degenerate particle filter evolved into the standard

77

particle filter with simple importance resampling (PF-SIR) is being used (Dubinkina et al. 2011;

78

Goosse et al., 2010, Goosse et al., 2012a, b; Annan and Hargreaves 2012; Mathiot et al., 2013;

79

Mairesse et al., 2013). In the PF-SIR, the analysis is derived as a weighted sum across all ensemble

80

members. The weights are calculated based on the closeness of each member to the available

81

observations. A resampling step is performed on the ensembles before proceeding to the next cycle

82

where the members with trivial weights are removed and the ones with non-trivial weights are

83

duplicated to maintain the same ensemble size. One advantage of the particle filter, in theory, is

84

its capability to provide the optimal analysis whether the error distribution is Gaussian or not

85

(Leeuwen 2009). However, the particle filter in general suffers from the so-called “curse of

86

dimensionality”, or particle impoverishment: A phenomenon where the non-trivial weights tend

87

to collapse to only a few particles while majority of the particles degenerate to trivial weights.

88

Resampling after each analysis step is used to alleviate such degeneracy, but PF-SIR still requires

89

a considerable amount of particles because the ensemble size required for a successful filtering

90

increases exponentially with the problem size (Snyder et al., 2008). Considering the large degree

91

of freedom of geophysical applications and the limited computation resources, the “curse of

92

dimensionality” has been a severe problem. However, some recent studies seem to be successful

93

in applying the PF-SIR for PDA using affordable ensemble sizes (N~100) (Dubinkina et al. 2011;

94

Goosse et al. 2011a, b; Mathiot et al., 2013, Mairesse et al., 2013). These studies did not encounter

95

filter degeneracy because they reduced the degree of freedom by performing time average, spatial

96

filtering, such as the long-term climatological temperature averaged in a single hemisphere.

97

Particle filter has also been used to investigate the causes and mechanisms of some climate

98

variability in the past (Goosse et al. 2011b; Mathiot et al., 2013).

99

The EnKF (Evensen, 1994, 2003) is a modern DA method that can be modified to

100

accommodate the time-averaged observations in paleo context (Dirren and Hakim 2005; Huntley

101

and Hakim 2010; Steiger et al. 2014; Hakim et al. 2016; Steiger and Hakim 2016). In the EnKF,

102

the final analysis equation is a linear combination of observation information and model estimate

103

with the weight determined according to the uncertainties (Evensen, 1994, 2003; Kalnay, 2003).

104

In the case of Gaussian error, the EnKF can achieve the optimal analysis. This optimal analysis in

105

theory maximizes the likelihood of achiving such an posterior estimates and minimizes the error

106

between itself and the prior estimates and the observations. In general, when the error is non-

107

Gaussian, however, there is no guarantee that the EnKF can achieve the optimal analysis as the

108

PF-SIR can. Nevertheless, the EnKF can provide robust and consistent analysis with relatively

109

small ensemble size as has been shown in numerous studies (e.g. Evensen, 1994, 2003; Zhang et

110

al. 2007)

111

Our goal is to perform a systematic comparison of the PF-SIR and the EnKF for

112

paleoclimate field reconstruction. We adopted on-line DA approach for both PF-SIR and EnKF,

113

which provides better temporal consistency (Matsikaris et al. 2015). The two methods will be

114

performed in pseudo proxy experiments (PPEs) which provide a synthetic, controlled test bed

115

(Smerdon 2012). We found that the EnKF with 16 ensemble members outperforms the PF-SIR

116

using 48 members in both the local reconstruction and hemispheric reconstruction. The conclusion

117

is robust even when the ensemble size of the PF-SIR is doubled to 96. This paper is organized as

118

follows. In section 2, we introduce the model and proxy network, evaluation metrics, and the

119

details of the assimilation methods. Section 3 compares the assimilation of the two filters and

120

discuss the advantages and disadvantages of each filter. The conclusion and some general

121

discussions will be given in section 4.

122 123

2 Model, Methods and Experimental Design

124 125

2.1 Model description

126

The Fast Ocean Atmosphere Model (FOAM) is used in this study. FOAM is a fully coupled global

127

atmosphere and ocean model (Jacob 1997). The atmosphere component is based on NCAR CCM2

128

and CCM3 models with R15 resolution (40 latitudes× 48 longitudes) and 18 vertical layers. The

129

ocean component is based on Modular Ocean Model (MOM) created by Geophysical Fluid

130

Dynamics Laboratory (GFDL), It has a regular 128x128 polar grid and 24 vertical levels. The

131

quality of the simulated climate compares well with higher resolution models. FOAM has been

132

used for various paleoclimate studies especially in Holocene (Liu et al. 2000; Liu et al. 2004; Liu

133

et al. 2006; Liu et al. 2007). It has also been used to test DA techniques and algorithms successfully

134

(Liu et al. 2014a; Liu et al. 2014b; Lu et al. 2015 part 2; Lu et al. 2016).

135 136

2.2 EAKF and its implementation.

137

Ensemble Adjustment Kalman Filter (EAKF) is used in FOAM. The EAKF is a variant of

138

EnKF and is a deterministic filter that does not require the perturbation of observations as EnKF

139

does (Anderson 2001, 2003). A brief introduction of the EnKF is provided below, which also

140

applies to the EAKF. For full comprehensive description, please refer to Anderson 2001, 2003 and

141

Kalnaey 2003.

142

The EnKF solves for the best least-square estimate of the system state given model

143

simulations and observations. Model simulations serve as a prior estimate (background) of the true

144

state, which is then updated with the observational information to solve for the posterior estimate

145

(analysis). The final analysis represents the most likely state of the climate system given model

146

estimates and observations. This process expressed in Bayesian framework is:

147

P(𝑿 | 𝒚) =

𝑷(𝒚|𝑿)𝑷(𝑿) 𝑷(𝒚)

(1)

148

Where 𝑷(𝑿) represents the prior distribution of the state X, P(𝒚 | 𝑿) represents the likelihood of

149

obtaining observation 𝒚 given model estimate 𝑿. 𝑷(𝒚) is a normalization term. P(𝑿 | 𝒚) is the

150

posterior distribution after considering observation information. The observations and the prior

151

estimate of the state are assumed to be unbiased with a Gaussian distribution. The model estimate

152

is characterized by its mean ̅̅̅ 𝑥 𝑏 and covariance matrix 𝑃𝑏 , where 𝑥 𝑏 is the prior state vector with

153

the dimension of the state size. The model estimate of the observation 𝑦 is then extracted from the

154

̅̅̅𝑏 ), where ℋ is the observation operator which projects the model space background ̅̅̅ 𝑥 𝑏 as ℋ(𝑥

155

into the observation space. The final analysis ̅̅̅ 𝑥 𝑎 is derived as a weighted sum of the observation

156

𝒚 and the model background ̅̅̅ 𝑥 𝑏 , and the weights are inversely proportional to their relative

157

uncertainty/error. The classic Kalman Filter update equations are:

158

̅̅̅𝑏 )) ̅̅̅ 𝑥 𝑎 = ̅̅̅ 𝑥 𝑏 + 𝐾(𝑦 − ℋ(𝑥

(2)

159

𝐾 = 𝑃𝑏 ℋ 𝑇 (ℋ𝑃𝑏 ℋ 𝑇 + 𝑅)−1

(3)

160

̅̅̅𝑏 )) of a grid point from a From the analysis equation, it can be see that the correction 𝐾(𝑦 − ℋ(𝑥

161

non-local observation is determined by the covariance between these two locations 𝑃𝑏 ℋ 𝑇 in the

162

gain matrix equation (3). If these two highly co-vary, the value of K is larger and hence it receives

163

a heavier correction, and vice versa. Therefore the EnKF and its variants are a covariance-based

164

method: as long as there is covariance, there is correction. In paleoclimatology, all proxy

165

observations contain time-averaged information, such as annual mean and seasonal mean

166

temperature, instead of instantaneous values like those of modern instrumental observations. The

167

algorithm is modified based on Hakim and Tardif (2005) to assimilate the time-averaged

168

observations. For example, if the observation 𝑦 in equation (2) represents seasonal mean

169

temperature, the corresponding model background ̅̅̅ 𝑥 𝑏 is also the seasonally averaged temperature.

170

Note that the covariance matrix 𝑃𝑏 is calculated from the time-averaged ensemble, therefore it

171

captures the low-frequency variability since the high-frequency variability has been smoothed out

172

during the time average.

173

To remove the long-distance spurious correlations caused by a limited ensemble size, a

174

covariance localization is implemented using the fifth order function of Gaspari and Cohn (1999).

175

Localization is an important technique used in the EnKF, which determines the influence radius

176

of the observation at a given location (Houtekamer and Mitchell 2001; Hamill and Whitaker 2001).

177

In a dense observation network, the localization radius is usually chosen to be small since most

178

model grids have a local observation constraint (Liu et al. 2014; Lu et al. 2014b, 2016). In a

179

spatially sparse observation network such as in the paleo context, a broader localization scale is

180

neccessary to extend the observation impact to larger regions. The localization scale in our EAKF

181

experiment is roughly 16 grid points in the ocean, which gives a distance of 5000 km on the equator,

182

and decreases poleward with the cosine of latitude. We will use “5000 km” to denote the EAKF

183

experiment with a localization scale of 16 grid points. Covariance inflation is also applied using

184

the relaxation method by Zhang et al. 2004 and the relaxation factor of 0.5. Further sensitivity tests

185

show that the performance is not sensitive to this parameter.. For the analysis cycle, we use an

186

“on-line” approach (with cycling).i.e. the analysis of the current step is used as the initial condition

187

of the next cycle. This EAKF method has been used to develop the first ensemble-based coupled

188

data assimilation (CDA) system in a fully coupled general circulation model (CGCM) (Zhang et

189

al.,2007). It has been implemented in FOAM to complete the first ensemble-based parameter

190

estimation experiment in a CGCM (Liu et al., 2014a, b).

191

2.3 Particle filter and its implementations

192

A standard particle filter with simple importance resampling (Gordon et al. 1993; Doucet et al.

193

2001; Leeuwen 2009), denoted as PF-SIR, is adopted in this work. The theoretical frameworkof

194

PF-SIR is also Bayes theorem as in equation (1). Here the particles refer to ensemble members,

195

and they provide a Monte Carlo approximation of the model distributions. For example,

196

given N particles, the prior distribution 𝑷(𝑿) can be constructed as a sum of 𝑁 delta functions

197

centered at each particle:

198

𝑃(𝑥) =

1 𝑁

∑𝑁 𝑖=1 𝛿(𝑥 − 𝑥𝑖 )

(4)

199

Given the prior distribution 𝑷(𝑿) and the likelihood 𝑷(𝒚|𝑿), the posterior density P (𝑿 | 𝒚) can

200

be approximated as:

201

𝑃(𝑥|𝑦) = ∑𝑁 𝑖=1 𝑤𝑖 𝛿(𝑥 − 𝑥𝑖 )

202

𝑤𝑖 =

with

𝑃(𝑦|𝑥𝑖 ) ∑𝑁 𝑗=1 𝑃(𝑦|𝑥𝑖 )

(5)

(6)

203

where 𝒘𝒊 is called the weight of the particle and it is a normalized likelihood. The likelihood

204

𝑃(𝑦|𝑥𝑖 ) is often taken to be Gaussian:

205

𝑃(𝑦|𝑥𝑖 ) = exp 𝑒 −

(ℋ(𝑥𝑖 )−𝑦)2 2𝐶

(7)

206

Where C is the error matrix and ℋ is the observation projector. Note that the likelihood of each

207

particle is a product of Gaussian weights at every observation location. Hence if the model estimate

208

differs significantly from the observation at only a few locations, the final weight will be reduced

209

to zero even if the majority of locations show a good match between the model and observation.

210

To mitigate the effects of the trivial weights at some locations on the final weight of the particle,

211

we add inflation to the weight calculated at every observation location before taken the product of

212

them. With the posterior distribution and weights, the final analysis will be:

213

214

̅̅̅̅̅̅ 𝑓(𝑥) = ∫ 𝑓(𝑥)𝑃(𝑥|𝑦)𝑑𝑥 ≈ ∑𝑁 𝑗=1 𝑤𝑖 𝑓(𝑥𝑖 )

(8)

215

Equations (4) through (7) construct the procedure of particle filter of sequential importance

216

sampling. The proposal density in this case is chosen to be the prior probability density and the

217

weights of the prior particles are equal (Doucet et al. 2001). This equal prior weight assumption is

218

achieved by resampling the particles at every filtering time step. Therefore, after importance

219

weights are calculated in equations (5) through (6), the particles are resampled according to their

220

importance weights. The particles with weights less than 𝑁 are dropped and particles with large

221

weights are duplicated. Though there are several resampling methods, we apply a systematic

222

resampling scheme (Hol et al. 2006) in this paper. After resampling, a small perturbation is added

223

to the identical particles. The identitical particles share the same ocean state, but differ in

224

atmospheric fields. These different atmospheric fields act as the perturbations. These new model

225

states will be used as initial conditions for the next cycle. In our study, about half of the 48 particles

226

end up having trivial weights and get removed at every analysis step, they are then replaced by the

227

duplications of the remaing particles.

1

228

Again, note that we are assimilating observations that contain time-averaged information,

229

hence the background is temporally averaged before calculating the importance weights. This

230

time-averaging also help to reduce the number of degree of freedom (Dubinkina et al. 2011). The

231

assimilation procedure is essentially the same as in the previous studies (Dubinkina et al. 2011;

232

Goosse et al. 2012a, b) except that no spatial filtering is applied. Both our PF-SIR and the PF-SIRs

233

in the previous studies referenced above do not have the localization implemented (The more

234

recently developed techniques in Poterjoy 2016 and Lee and Majda 2016).

235

2.4 Pseudo proxy network and experiments

236

We use pseudo seasonal mean sea surface temperature (SST) observations to conduct the

237

PPEs. The pseudo proxy locations are chosen from Marcott et al. 2013, which provides 47

238

observations globally (Figure 1a) with a higher density in the northern hemisphere. The pseudo

239

proxy time series are generated by adding Gaussian noise with a standard deviation of 0.5 °𝐶 to a

240

specific model control run representing the actual state or “truth”. The Gaussian noise mimics the

241

errors associated with the proxy observation. The performance of each DA method is measured

242

using the root-mean-square error (RMSE) and correlation coefficient between the analysis and the

243

truth.

1

𝑇 2 𝑅𝑀𝑆𝐸𝑖 =√𝑛𝑡 ∑𝑡[𝑋𝑖,𝑡 − 𝑋𝑖,𝑡 ]

244

(9)

245

𝑇 where 𝑖 represents the location, 𝑡 represents the time, 𝑋𝑖,𝑡 is the truth and 𝑋𝑖,𝑡 is the analysis. In all

246

experiments, we use 16 members for EAKF and 48 for PF-SIR. We did not use the same ensemble

247

size because EAKF shows decent performance even using a relatively small number of ensemble

248

members (Lu et al. 2014b, 2016). Our results will show that even with only 16 members, EAKF

249

still significantly outperforms PF-SIR with 48 members.

250

A 200-year-long experiment is carried out for both EAKF and PF-SIR assimilating

251

seasonal SST pseudo observations. Two EAKF sensitivity experiments were also carried out using

252

smaller localization scale, 2000 km and 500km, to illustrate the importance of localization radius

253

to the performance of the EAKF. Furthermore, to test the sensitivity of ensemble size, we also

254

carried out a PF-SIR experiment of 96 members and a EAKF-5000km experiment of 32 members.

255

These two ensemble size sensitivity tests are only 50-year long because of the expensive

256

computational costs.

257

3 Results

258

The performance of the two filters was mainly assessed by computing the correlation on

259

both local scale and hemispheric scale between the analysis SST and the truth. We also investigate

260

the ability of each filter in capturingto capture the modes of variability by performing empirical

261

orthogonal function (EOF) analysis on both the simulation and the truth.

262

3.1 Local reconstruction skills

263

Figure 1 displays the analysis quality of SST characterized by RMSE and correlation

264

coefficient for the PF-SIR and the EAKF of 5000 km localization scale. The RMSE is calculated

265

from the differences between seasonally-averaged ensemble mean analysis and the “truth” at each

266

grid point and normalized by its corresponding natural variability from Figure 1a. The local

267

reconstruction skill of the EAKF is significantly better than that of the PF-SIR at both the proxy

268

location and the nearby regions. In the North Atlantic where the observations are dense, the

269

correlation with truth is above 0.6 basin-wide for the EAKF (Figure 1e) but below 0.6 for the PF-

270

SIR (Figure 1c); and the RMSE ranges from 0.4 to 0.6 for the EAKF (Figure 1d) but above 0.8 for

271

the PF-SIR (Figure 1b). In regions that have observation clusters such as the south of Iceland and

272

the east of Newfoundland, the correlation coefficient for the EAKF reaches 0.85. In the tropical

273

eastern Pacific where there is only one observation in the far east, the EAKF still maintains a

274

correlation of around 0.4 and RMSE around 0.9 (Figure 1e and d), but the PF-SIR’s reconstruction

275

skill is fairly weak (Figure 1b and c).

276

The spatial reconstruction skill of the EAKF greatly depends on the localization scale.

277

Recall that localization determines how far can the observation at a given location influence its

278

nearby locations. In the bottom two rows of Figure 1, we show the results of the EAKF with

279

decreasing localization scales of 2000 km (Figure 1f, g) and 500 km (Figure 1h, i). When the

280

localization scale is reduced, the area of higher correlation and lower RMSE also shrink towards

281

the observation location. When the localization scale is reduced to 500 km, well-constrained

Formatted: Left, Indent: First line: 0.5", Widow/Orphan control

282

regions shrink to small areas tightly surrounding the proxy locations, as colored patches in

283

Figure 1d and 1e diminish to colored dots in Figure 1h and 1i.

284

In PDA, larger localization scale can extend the observational impact of the sparse proxy

285

network, especially in regions where large-scale variability exists and only a few proxies are

286

available. For example, there are basin-wide covarying patterns, such as Pacific Decadal

287

Oscillation (PDO) in the north Pacific and El Nino-Southern Oscillation (ENSO) in the tropical

288

Pacific. One observation along the west coast of North America (Figure 1a) is able to partially

289

constrain the evolution of PDO in the entire basin. This is also demonstrated in section 3.3 in terms

290

of modes of variability. When localization radius is small, the reconstruction skill is lost away

291

from the observation sites.

292 293

3.2 Hemispheric averaged SST time series

294 295

Figure 2 shows the hemispheric averaged SST time series, derived as area weighted mean

296

of the analysis. For the northern hemisphere (NH) where most of the observations are located and

297

where the ocean area is not overwhelmingly dominant (~60% coverage) (Figure 2d), the EAKF

298

with 5000 km localization scale (red line) and PF-SIR (blue line) have a correlation with the true

299

state (black line) of 0.85 and 0.67, respectively. They both capture the decadal scale variability

300

very well, except that during the year 30 to 80 the PF-SIR starts to suffer from particle

301

impoverishment where it cannot select good members that match the observations out of its 48

302

members. This mismatch greatly diminishes the final correlation coefficient calculated between

303

PF-SIR and the truth. To take such situations into account, we calculate the correlation coefficient

304

between the truth curve and the other three curves in Figure 2 a, d, and g at different time segments

305

(Figure 3), such as 20 years, 40 years and 100 years. These segments are sampled from the 200-

306

year simulation. Thus each dot in Figure 3 represents the correlation coefficient of a sample. We

307

can see that the EAKF- 5000km has higher correlation coefficient than the PF-SIR at 20 years, 40

308

years, 100 years and 200 years sample length (Figure 3a). The distribution of the dots is more

309

compact for the EAKF-5000km than the PF-SIR, which suggests that the performance of the

310

EAKF-5000km is more stable than the PF-SIR. Similar results can also be found in Figure 3b-c.

311

In addition, despite that PF-SIR captures the general variability in the northern hemisphere, there

312

is a small phase shift between the PF-SIR and the truth. This again reflects the nature of the PF-

313

SIR in that the analysis is reconstructed from existing ensemble members that are closest to the

314

observation and the analysis will deviate from the truth if all the members are far away from the

315

observations. In contrast, the EAKF corrects each member using the observation increment so that

316

every member in the ensemble will be “dragged” towards the true state. Hence the final analysis

317

has better synchronization with the truth time series compared to the PF-SIR. For the southern

318

hemisphere (SH), both methods failed to track the true state due to very limited observations and

319

the large area of ocean (~80% coverage) (Figure 2g). On the global scale (Figure 2a), the

320

correlation with truth for the global averaged SST is 0.61 for the EAKF and 0.24 for the PF-SIR.

321

The low correlation coefficient of the PF-SIR is further reduced by the poor performance from the

322

SH. Therefore, even though the PF-SIR is deficient at local reconstruction as shown in Figure 1b

323

and c, it seems to be able to capture some hemispheric-scale variability if the ensemble size is

324

sufficiently large. However, even with 3 times the computational expense, the PF-SIR with 48

325

members still underperforms the EAKF with 16 members in the reconstruction of hemispheric-

326

scale SST.

327

The uncertainty of the final analysis is also shown in Fig.2, as represented by its ensemble

328

members. The second column is for the EAKF with 5000 km localization radius (the pink curves

329

as the 16 ensemble members)PF-SIR (light blue curves as the 48 ensemble members) and the third

330

column is for the PF-SIR (light blue curves as the 48 ensemble members) EAKF with 5000 km

331

localization radius (the pink curves as the 16 ensemble members). In general, the PF-SIR tends to

332

underestimate the uncertainty of the model forecasts, an indication of the problem of filter

333

degeneracy. All the 48 ensemble members cluster tightly at the hemispheric and global scales and

334

the spread is not large enough to encompass the truth time series. This is because the posterior

335

ensemble are heavily bounded by the prior ensemble: it is impossible to generate the posterior

336

ensemble outside the span of the prior ensemble even when the observation is outside the span.

337

For the EAKF, the truth series always lies within the uncertainty of the 16 ensemble members. In

338

the NH (Figure 2f), abundant observations lead to stronger constraint of the analysis and hence

339

smaller uncertainty, while in the SH (Figure 2i), weak observational constraint causes large spread

340

among the 16 ensemble members. This suggests that the EAKF provides a better representation of

341

the uncertainty with only 16 members than the PF-SIR with 48 members.

342 343

3.3 Modes of climate variability

344 345

In this section, we investigate the ability of the two methods in capturing the modes of

346

climate variability. Empirical Orthogonal Function (EOF) analysis is performed on the four PDA

347

analyses (PF-SIR and three EAKF experiments with different localization radius) as well as the

348

true state on different regions. The EOF patterns and principle components (PCs) of the first mode

349

of variability are shown in Figure 4 and Figure 6-8. In the observation-dense North Atlantic (Right

350

column in Figure 4), both the PF-SIR and the EAKF-5000 km capture the 1st EOF pattern that is

351

almost identical to the truth in both pattern and strength, and the pattern correlation is close to 1

352

for both experiments (Table 1). The corresponding PCs of these two methods show similar decadal

353

variability to the truth. The temporal correlation coefficients between the PC time series of each

354

filter and that of the truth is as high as 0.78 for EAKF-5000km and 0.67 for PF-SIR (Table 1).

355

However, the explained variance of the first mode is much higher in the PF-SIR (25.9%) and

356

EAKF-5000km (24.7%) than that in truth (18.8%), which suggests that both filters have trouble

357

constraining the precise variability given so limited observational information. As the localization

358

radius in the EAKF is decreased, the spatial pattern gradually becomes weaker (Figure 4f, h and j)

359

and the temporal correlation coefficient also decreases (Table 1). Consistent with the grid point

360

analysis and hemispheric analysis, a large localization scale is necessary even in relatively dense

361

observation area.

362

In the North Pacific, PF-SIR (Figure 4c) captured clearly a PDO pattern similar to the truth

363

(Figure 4a).The correlation coefficient between the corresponding PC time series and the truth in

364

the northern hemisphere is 0.23 for PF-SIR and 0.57 for EAKF- 5000 km (Table 1). Note that there

365

are very few observations in the North Pacific: one around the Oregon coast and a few around east

366

China Sea (Figure 1a). The vast majority of the ocean basin remains unobserved. Nevertheless,

367

both PF-SIR and EAKF- 5000 km can capture the PDO spatial pattern and constrain the temporal

368

variability. However, the temporal variability of the PF-SIR is much less accurate than that of the

369

EAKF- 5000 km. The ability of the EAKF to constrain the entire basin using very few observations

370

is due to two factors: the existence of a large-scale co-varying pattern, namely PDO, and the

371

existence of observations in the critical region, in this case, around the Oregon coast. The control

372

of this basin-wide co-variability through the critical observation location is reflected by the one-

Formatted: Indent: First line: 0.5"

373

point correlation map in Figure 5a. The observation in the crucial region therefore provides

374

information about the phase of the large-scale co-variability. With this information, the PF-SIR

375

can select members that have the correct phase and the EAKF can nudge the impact area towards

376

the targeted phases. In this case, EAKF requires a large localization radius such that the entire co-

377

varied area can be corrected. As can be seen from Figure 4e, g and i, the spatial pattern gradually

378

fades away as the localization scale is decreased, as does the temporal correlation with truth (Table

379

1).

380

As is the case in the North Pacific, the tropical Pacific also has large-scale co-variability

381

with the observation location (Figure 5b) and only one observation located in the far eastern

382

portion of the basin (Figure 1a). Like north Pacific, both PF-SIR and EAKF 5000 km captured the

383

pattern of ENSO with pattern correlation coefficient close to 1, although the magnitude is weaker

384

than truth. The EAKF-5000km shows better temporal variability, as the correlation in the first PC

385

is 0.43 compared to 0.25 for the PF-SIR. The weaker spatial pattern but better temporal variability

386

in the tropics and North Pacific for the EAKF compared to the PF-SIR is due to the fundamental

387

differences of the two filters. The analysis of EAKF is ensemble mean output, thus when

388

observational constraint is weak, the ensemble mean tends to average out the internal variability

389

and result in a weaker spatial pattern. Also, recall that in the EAKF the physical field in each

390

member is modified by the observation innovation. Hence given the phase information of the large-

391

scale pattern from the key observation, the EAKF will directly correct all members and drag them

392

towards the correct phase. The PF-SIR does not have this feature because the analysis of the PF-

393

SIR is a weighted sum across all members (The physical field in each member is not modified),

394

and the weight attached to each member is based on observations from all locations. Therefore,

395

the local constraint of a given observation will be mitigated by observations from other locations.

396

Thus, the temporal constraint of these key observations on their neighboring regions is weaker in

397

the PF-SIR than that in the EAKF. As localization radius is decreased in the two additional

398

experiments of the EAKF, the first EOF pattern gets weaker in the tropics and the PC correlation

399

with truth also decreases. When localization radius is decreased to 500 km, the EAKF barely has

400

any constraint of the large-scale variability in this area (Figure 8 and Table 1).

401

In the SH where there are very few observations, both methods failed to reconstruct the

402

leading modes of similar patterns to the truth (Figure 7). The spatial field of the PF-SIR bears some

403

similarities to the truth in the south Pacific, but the temporal variability of this pattern has no

404

correlation with that of truth at all. The three EAKF experiments capture a pattern that co-varies

405

with the truth with a correlation coefficient around 0.3 (Table 1), but these patterns do not

406

resemble the truth that much. The EAKF- 5000 km maintains some performance near observation

407

locations such as around South Atlantic and the west coast of Chile and New Zealand (green

408

patches in Figure 1e). But these limited observations are far from enough to constrain the entire

409

southern hemisphere considering that there is no large-scale co-variability in the southern

410

hemisphere as in the northern hemisphere (Figure 5c and d). Hence it is not surprising that both

411

methods perform poorly in the southern hemisphere (Table 1).

412 413

3.4 Reconstruction skills on other variables

414

From the reconstructed SST analysis, we see that the EAKF-5000km has better

415

performance than PF-SIR is both the local reconstruction and hemispherically averaged

416

reconstruction. In this section, we will take look at the spatial reconstruction skills on other

417

variables, namely air surface temperature (AST), 500hPa geopotential height (Z) and precipitation

418

(PRECP) for both filters (Figure 9). Since RMSE ratio map shows similar skills to spatial

419

correlation map, we only show correlation map for these variables. The correlation coefficients are

420

calculated over the entire 200-year simulation after smoothing the time series at each grid point to

421

remove the high-frequency noise in the atmosphere. The AST reconstructions (Figure 9a-b) are

422

similar to the SST reconstruction (Figure 1c and e) over the ocean area, but have a slightly larger

423

impact region away from the observation location (Figure 9a-b). Over land, the reconstruction

424

skills mainly exist in coastal regions where there are observations in the nearby ocean, such as

425

Eastern North America, Greenland, North Africa and East Asia. Again, EAKF outperforms the

426

PF-SIR by showing larger area of high correlation coefficients especially in the Pacific ocean and

427

the south Atlantic ocean. For 500hPa geopotential height, the patterns are similar to AST (Figure

428

9c-d). Both PF-SIR and EAKF have poor reconstruction skills for precipitation. This is because

429

that precipitation is not only influenced by temperature, but also strongly influenced by the

430

moisture distribution and the wind field. Therefore it is not surprising to see such poor

431

reconstruction skills on precipitation from both filters.

432 433

3.5 sensitivity to ensemble size and localization scale

434 435

Considering a sparse observation network in PDA, the large localization radius is crucial

436

for EAKF to extend the observation impact and hence to constrain large scale variability as seen

437

from previous sections. This leads to the question on the optimal localization radius for such PDA

438

application. To find such optimal localization radius, we further increased 5000km to 10,000km

439

in the EAKF. This is almost equivalent to that each observation updates the entire globe. We found

440

that in terms of local reconstruction skill, North Pacific and tropical Pacific have stronger PDO

441

and ENSO pattern compared to the EAKF-5000km, but except these regions, the patterns in other

442

regions are almost identical (Figure not shown). In terms of hemispheric averaged SST time series,

443

the correlation for the EAKF-10000km is even slightly smaller than the EAKF-5000km. This is

444

because in regions that do not have the large-scale co-variability, a large localization radius allows

445

sampling error to come into play, which eventually deteriorates the hemispheric-scale performance.

446

Therefore, although large localization scale is necessary for PDA applications, it is not “the larger,

447

the better”. This suggests that the optimal localization radius for PDA applications is not a globally

448

uniform value, but a spatially-varying parameter. Zhen and Zhang 2014 has proposed a variational

449

approach to adaptively determines the optimum radius of influence for ensemble covariance

450

localization. However, the spatially-varied localization parameter is beyond the scope of this paper,

451

but is worthy of future investigation.

452

Ensemble size is an important factor for ensemble filters, especially for the PF-SIR which

453

requires an ensemble size that increases exponentially with the dimension of the system (Snyder

454

et al. 2008), and previous studies adopting on-line PF-SIR uses 96 ensemble members (Dubinkina

455

et al. 2011; Goosse et al. 2012). We double the ensemble size of the EAKF-5000km and PF-SIR

456

and run the experiment for 50 years to see if this will greatly improve the performance of the two

457

filters and potentially change the relative performance. For the EAKF-5000km, doubling the

458

ensemble size yield slight improvement in local skill (Figure 10b-c) compared to previous case

459

using only 16 members (Figure 1d-e). One noticeable improvement is in the tropical Pacific. For

460

the northern hemispheric SST time series, the correlation with truth has increased from 0.79 to

461

0.91. The global mean temperature correlation has increased from 0.54 to 0.72 (Figure 2). For the

462

PF-SIR, there is noticeable improvement in the local reconstruction scale (Figure 10d-e) such as

463

in the North Pacific and South Pacific. But for hemispheric averaged SST, there is no significant

464

improvement. EAKF-5000km still outperforms the PF-SIR when ensemble size is doubled for

465

both filters. In fact, EAKF-5000km with only 16 members still significantly outperforms the PF-

466

SIR with 96 members in both local skill and hemispheric SST time series reconstruction.

467 468

4 Summary and discussion

469 470

Particle filter with simple importance resampling and ensemble Kalman filter are emerging

471

as two popular ensemble-based data assimilation methods for paleoclimate studies. Here, using a

472

CGCM (FOAM), we systematically compare the performance of the two filters in assimilating

473

seasonally averaged observations of SST in PPEs and provide some insights into the choice for

474

PDA methodology. It is found that EnKF has overall better performance than the PF-SIR using

475

only one third the number of ensemble members. On local scales, EnKF shows good reconstruction

476

skill in the northern hemisphere where observations are relatively abundant, as well as in regions

477

that have few observations but has large-scale variability of coherent pattern, such as the tropical

478

Pacific and the North Pacific. For the PF-SIR, the correlation with truth is much smaller compared

479

to the EnKF and only the north Atlantic region has non-trivia correlation coefficient. In the

480

hemispheric reconstruction, both methods can capture the decadal variability in the data-rich

481

northern hemisphere and lose skill in the data-sparse southern hemisphere. In general, the EnKF

482

has higher correlation with the actual time series than the PF-SIR. PF-SIR is also subject to particle

483

impoverishment where all members end up far from the observations and hence few particles can

484

be selected out of its 48 ensemble members. When such degeneracy occurs, the reconstruction will

485

deviate from the truth substantially as shown in the year 30 - 80 in the 200-year-long experiment

486

(Figure 2). Doubling the ensemble size of the PF-SIR improve the local reconstruction, but the

487

overall reconstruction skill is still smaller than the EnKF with only 16 members. Hence we

488

conclude that the EnKF is more consistent and stable with much cheaper computation cost

489

compared to the PF-SIR.

490

The success of the EAKF over the PF-SIR in this PPE is attributed to the fact that a)

491

Gaussian assumption is appropriate for this application. b) The EAKF is less sensitive to sampling

492

errors than PF-SIR due to different natures of the two filters. 16 members is enough to capture the

493

covariance used in equation (3), but 48 (even 96) members still underrepresent the state space of

494

such high-dimensional system. This underrepresentation is manifested as the bounding of the

495

posterior particles of prior ensemble, since it is impossible to generate posterior samples outside

496

the span of the prior ensemble even when the observation is far outside of the prior ensemble.

497

Therefore a sufficient ensemble size is critical to the PF-SIR, and this “sufficient” ensemble size

498

is much greater than that required by the EAKF. c)Note that one important practice of applying

499

EAKF to PDA is the choice of a large localization radius since there There exists is large-scale co-

500

variability on slow time scales., which This practice benefits the covariance-based filter

501

substantially. As the localization radius of the EnKF is decreased in the two addition experiments,

502

the performances on both the local scale and hemispheric scale are diminished. This localization

503

radius effect is most obvious in regions that have large-scale co-variability such as tropical Pacific

504

and the North Pacific. One observation in these regions is able to constrain the entire basin given

505

a larger observation impact radius. Three EAKF experiments with different localization scales has

506

clearly illustrated this point, but further study can be conducted to apply different localization

507

radius at different grid points: regions that have large-scale co-variability are assigned larger

508

localization radius, while regions that do not have such feature or have relatively dense

509

observations are assigned smaller localization scale. This localized localization radius network

510

will exploit the sparse observation information and at the same time avoid the long-distance

511

spurious correlation, which will be very suitable to PDA applications.

512

These results raise the question of which filter should be preferred in the future for

513

paleoclimate studies. In our experiments, we show that the EnKF with only 16 members

514

outperforms the PF-SIR with 48 (even 96) members in almost every way. In theory particle filter

515

allows non-Gaussianity, which is an advantage over the EnKF. The challenge for particle filter is,

516

however, the “curse of dimensionality”. To better apply particle filter for paleoclimate studies,

517

more techniques such as localized and clustered particle filter (Poterjoy 2016; Lee and Majda 2016)

518

should be considered instead of using just the particle filter with simple importance resampling.

519 520

Acknowledgement

521 522

We gratefully appreciate the help of Zachary R. Hansen in editing the manuscript. This work is

523

supported by a grant from NUIST, Chinese NSF41630527 and NSF P2C2 program. The data used

524

in this paper is stored in UCAR HPSS system and is available upon request from

525

[email protected].

526

527

References

528

Anderson, J. L., (2001), An Ensemble Adjustment Kalman Filter for Data Assimilation. Mon.

529

Weather Rev., 129, 2884–2903, doi:10.1175/15200493(2001)1292.0.CO;

530

2.

531 532

533 534

Anderson, J. L., (2003), A local least squares framework for ensemble filtering. Mon. Weather Rev., 131, 634–642, doi:10.1175/1520-0493(2003)1312.0.CO;2.

Annan, J. D., and J. C. Hargreaves, (2012), Identification of climatic state with limited proxy data. Clim. Past, 8, 1141–1151, doi:10.5194/cp-8-1141-2012.

535

Braconnot, P., S. P. Harrison, M. Kageyama, P. J. Bartlein, V. Masson-Delmotte, A. Abe-Ouchi,

536

B. Otto-Bliesner, and Y. Zhao, (2012), Evaluation of climate models using palaeoclimatic

537

data. Nat. Clim. Chang., 2, 417–424, doi:10.1038/nclimate1456.

538 539

Barkmeijer, J., Iversen, T. and Palmer, T. N. (2003), Forcing singular vectors and other sensitive model structures. Q.J.R. Meteorol. Soc., 129: 2401–2423. doi:10.1256/qj.02.126

540

Bhend, J., J. Franke, D. Folini, M. Wild, and S. Brönnimann, (2012), An ensemble-based approach

541

to climate reconstructions, Clim. Past, 8(3), 963–976, doi:10.5194/cp-8-963-2012.

542

Crespin, E., H. Goosse, T. Fichefet, and M. E. Mann, (2009), The 15th century Arctic warming

543

in coupled model simulations with data assimilation. Clim. Past, 5, 389–401,

544

doi:10.5194/cp-5-389-2009.

545 546 547

Doucet, A., N. de Freitas, and N. Gordon, (2001), Sequential MonteCarlo Methods in Practice. Springer-Verlag, 581 pp.

548 549

Dirren, S., and G. Hakim, (2005), Toward the assimilation of timeaveraged observations. Geophys. Res. Lett., 32, L04804, doi:10.1029/2004GL021444.

550

Dubinkina S, Goosse H, Sallaz-Damaz Y, Crespin E, Crucifix M, (2011) Testing a particle filter

551

to reconstruct climate changes over the past centuries. Int J Bifurc Chaos 21(12):3611–3618

552

Evensen, G., (1994), Sequential data assimilation with a nonlinear quasi-geostrophic model using

553

Monte Carlo methods to forecast error statistics. J. Geophys. Res., 99, 10143,

554

doi:10.1029/94JC00572.

555 556

557 558

Evensen, G., (2003), The Ensemble Kalman Filter : theoretical formulation and practical implementation. 343–367, doi:10.1007/s10236-003-0036-9.

Gordon N J, Salmond D J, Smith A F M. (1993), Novel approach to nonlinear/non-Gaussian Bayesian state estimation. IEE P F Radar Sign Proc,140: 107–113

559

Goosse, H., H. Renssen, A. Timmermann, R. Bradley, and M. Mann (2006a), Using paleoclimate

560

proxy-data to select optimal realisations in an ensemble of simulations of the climate of the

561

past millennium, Clim. Dyn., 27, 165–184, doi:10.1007/s00382-006-0128-6.

562

Goosse, H, O. Arzel, J. Luterbacher, M. E. Mann, H. Renssen (2006b), The origin of the European

563

"Medieval Warm Period". Climate of the Past Discussions, European Geosciences Union

564

(EGU), 2 (3), pp.285-314.

565

Goosse, H, E. Crespin, A. de Montety, M.E. Mann, H. Ressen and A. Timmermann (2010a),

566

Reconstructing surface temperature changes over the past 600 years using climate model

567

simulations

with

data

568

doi:10.1029/2009JD012737.

assimilation,

J.

Geophys.

Res.,

115,

D09108,

569

Goosse, H., Lefebvre, W., de Montety, A., Crespin, E., Orsi, A., (2009). Consistent past half-

570

century trends in the atmosphere, the sea ice and the ocean at high southern latitudes. Climate

571

Dynamics 33, 999e1016.

572

Goosse, H., J. Guiot, M. E. Mann, S. Dubinkina, and Y. SallazDamaz, (20121a): The medieval

573

climate anomaly in Europe: Comparison of the summer and annual mean signals in two

574

reconstructions and in simulations with data assimilation. Global Planet. Change,

575

doi:10.1016/j.gloplacha.2011.07.002

576

Goosse, H., J.Guiot, M.E. Mann, S.Dubinkina, and Y. Sallaz-Damaz (20121b), The Medieval

577

Climate Anomaly in Europe in simulations with data assimilation. Quat. Int., 279–280, 173–

578

174, doi:10.1016/j.quaint.2012.08.246.

579 580

581 582

Hakim, G. J., J. Emile-geay, E. J. Steig, D. Noone, and M. David, (2016), The Last Millennium Climate Reanalysis Project : Framework and First Results. 98195, 1–56.

Huntley, H. S., and G. J. Hakim (2010), Assimilation of time-averaged observations in a quasigeostrophic atmospheric jet model, Clim. Dyn., 35(6), 995–1009.

583

Houtekamer, P. L., and H. L. Mitchell, (2001), A sequential ensemble {Kalman} filter for

584

atmospheric data assimilation. Mon. Weather Rev., 129, 123–137, doi:10.1175/1520-

585

0493(2001)1292.0.CO;2.

586

Hamill, T. M., J. S. Whitaker, and C. Snyder, (2001): Distance-Dependent Filtering of Background

587

Error Covariance Estimates in an Ensemble Kalman Filter. Mon. Weather Rev., 129, 2776–

588

2790, doi:10.1175/1520-0493(2001)1292.0.CO;2.

589

Hol, J. D., T. B.Schon, and F.Gustafsson (2006), On resampling algorithms for particle filters,

590

paper presented at Nonlinear Statistical Signal Processing IEEE Workshop, Linkoping,

591

Sweden.

592

593

594 595

596 597

598 599 600

Jacob, R., (1997): Low Frequency Variability in a Simulated Atmosphere Ocean System.

University of Wisconsin-Madison, 159 pp.

Kalnay E. (2003), Atmospheric Modeling, Data Assimilation and Predictability. 4th ed. Cambridge: Cambridge University Press

Liu, Z., J. E. Kutzbach, and L. Wu (2000), Modeling climatic shifts of el Nino variability in the Holocence. Geophys. Res. Lett., 27, 2265–2268, doi:10.1029/2000GL011452.

Liu, Z., Harrison, S.P., Kutzbach, J., Otto-Bliesner, B., (2004). Global monsoons in the MidHolocene and oceanic feedback. Clim. Dyn. 22, 157–182. Liu, Z., Wang, Y., Gallimore, R., Notaro, M., Prentice, I.C., (2006). On the cause of abrupt

601

vegetation collapse in North Africa during the Holocene: climate variability vs. vegetation

602

feedback. Geophys. Res. Lett. 33, L22709. doi:10.1029/2006GL028062.

603

Liu, Z., Wang, Y., Gallimore, R., Gasse, F., Johnson, T., deMenocal, P., Adkins, J., Notaro, M.,

604

Prentice, I.C., Kutzbach, J., Jacob, R., Behling, P., Wang, L., Ong, E., (2007), Simulating

605

the transient evolution and abrupt change of Northern Africa atmosphere–ocean–terrestrial

606

607 608

609

ecosystem in the Holocene. Quaternary Science Reviews 26, 1818–1837.

Leeuwn, P., (2009), Particle Filtering in Geophysical Systems. Monthly Weather Reviews, 4089– 4114, doi:10.1175/2009MWR2835.1.

Liu, Y., Z. Liu, S. Zhang, X. Rong, R. Jacob, S. Wu, and F. Lu, (2014a), Ensemble-based

610

parameter estimation in a coupled GCM using the adaptive spatial average method. J.

611

Climate, 27, 4002–4014, doi:10.1175/JCLI-D-13-00091.1.

612

Liu, Y., Z. Liu, S. Zhang, R. Jacob, F. Lu, X. Rong, and S. Wu, (2014b), Ensemble-based

613

parameter estimation in a coupled general circulation model. J. Climate, 27, 7151–7162,

614

doi:10.1175/ JCLI-D-13-00406.1.

615 616

617

Lee, Y., and A. J. Majda, (2016), State estimation and prediction using clustered particle filters. Proc. Natl. Acad. Sci., 113, 201617398, doi:10.1073/pnas.1617398113.

Lu F, Liu Z, Zhang S, Liu Y, Jacob R, (2015), Strongly coupled data assimilation using leading

618

averaged coupled covariance (LACC). Part II: CGCM experiments*. Mon Weather Rev

619

143:4645–4659. doi:10.1175/MWR-D-15-0088.1

620

Lu, F., Liu, Z., Liu, Y., (2016). Understanding the control of extratropical

621

atmospheric variability on ENSO using a coupled data assimilation approach

622

doi:10.1007/s00382-016-3256-7

623

624

Marcott, S. a., J. D. Shakun, P. U. Clark, and A. C. Mix, (2013), A Reconstruction of Regional.

Science, 339, 1198–1201, doi:10.1126/science.1228026.

625

Mann, M. E., Z. Zhang, M. K. Hughes, R. S. Bradley, S. K. Miller, S. Rutherford, and F. Ni,

626

(2008), Proxy-based reconstructions of hemispheric and global surface temperature variations

627

over the past two millennia. Proc. Natl. Acad. Sci. U. S. A., 105, 13252–13257,

628

doi:10.1073/pnas.0805721105.

629 630

Mann, M. E., and Coauthors, (2009), Global signatures and dynamical origins of the Little Ice Age and Medieval Climate Anomaly. Science, 326, 1256–1260, doi:10.1126/science.117 7303.

631

Mathiot, P., H. Goosse, X. Crosta, B. Stenni, M. Braida, H. Renssen, and C. J. Van Meerbeeck,

632

(2013), Using data assimilation to investigate the causes of Southern BP. Clim. Past, 887–

633

901, doi:10.5194/cp-9-887-2013.

634

Mairesse, A., H. Goosse, P. Mathiot, H. Wanner, and S. Dubinkina, (2013), Investigating the

635

consistency between proxy-based reconstructions and climate models using data assimilation:

636

a mid-Holocene case study. Clim. Past, 5194, doi:10.5194/cp-9-2741-2013.

637 638

Matsikaris, A., M. Widmann, J. Jungclausn, (2015), On-line and off-line data assimilation in palaeoclimatology : a case study. Clim. Past, 11:81–93. doi:10.5194/cp-11-81-2015.

639

Otto-Bliesner, B. L., E. C. Brady, G. Clauzet, R. Tomas, S. Levis, and Z. Kothavala, (2006), Last

640

glacial maximum and Holocene climate in CCSM3. J. Clim., 19, 2526–2544,

641

doi:10.1175/JCLI3748.1.

642 643

Poterjoy, J., 2015: A localized particle filter for high-dimensional nonlinear systems. Mon. Weather Rev., 151020092806005, doi:10.1175/MWR-D-15-0163.1.

Formatted: Font: Italic

644

Snyder, C., T. Bengtsson, P. Bickel, and J. Anderson, (2008), Obstacles to High-Dimensional

645

Particle Filtering. Mon. Weather Rev., 136, 4629–4640, doi:10.1175/2008MWR2529.1.

646

Smerdon, J. E., (2012), Climate models as a test bed for climate reconstruction methods:

647

Pseudoproxy

648

doi:10.1002/wcc.149.

649 650

651 652

653

experiments.

Wiley

Interdiscip.

Rev.

Clim.

Chang.,

3,

63–77,

Steiger, N. J., G. J. Hakim, E. J. Steig, D. S. Battisti, and G. H. Roe (2014), Assimilation of timeaveraged pseudoproxies for climate reconstruction, J. Clim., 27(1), 426–441

Steiger, N., and G. Hakim, (2016), Multi-timescale data assimilation for atmosphere-ocean state estimates. Clim. Past, 12, 1375–1388, doi:10.5194/cp-12-1375-2016.

Tippett, M. K., J. L. Anderson, C. H. Bishop, T. M. Hamill, and J. S. Whitaker, (2003), Ensemble

654

Square

Root

Filters.

Mon.

Weather

655

0493(2003)1312.0.CO;2.

Rev.,

131,

1485–1490,

doi:10.1175/1520-

656

Von Storch H, Cubasch U, Gonzalez-Rouco J F, Jones J M, Voss R, Widmann M, Zorita E. ,

657

(2000), Combining paleoclimatic evidence and GCMs by means of data assimilation through

658

upscaling and nudging (DATUN). In: Proceedings of the 11th Symposium on Global Change

659

Studies. 28–31

660

Whitaker, J. S., and T. M. Hamill, (2002): Ensemble data assimilation without perturbed

661

observations. Mon. Wea. Rev., 130, 1913–1924, doi:10.1175/1520-0493(2002)130,1913:

662

EDAWPO.2.0.CO;2.

663

Widmann, M., H. Goosse, G. van der Schrier, R. Schnur, and J. Barkmeijer, (2010), Using data

664

assimilation to study extratropical Northern Hemisphere climate over the last millennium.

665

Clim. Past, 6, 627–644, doi:10.5194/cp-6-627-2010.

666

Zhang, F. Q., C. Snyder, and J. Z. Sun, 2004: Impacts of initial estimate and observation

667

availability on convective-scale data assimilation with an ensemble Kalman filter. Mon. Wea.

668

Rev., 132, 1238–1253,

669

Zhang, S., M. J. Harrison, a. Rosati, and a. Wittenberg, (2007), System Design and Evaluation of

670

Coupled Ensemble Data Assimilation for Global Oceanic Climate Studies. Mon. Weather

671

Rev., 135, 3541–3564, doi:10.1175/MWR3466.1.

672

Zhen, Y., and F. Zhang, 2014: A probabilistic approach to adaptive covariance localization for

673

serial

674

doi:https://doi.org/10.1175/MWR-D-13-00390.1

675

ensemble

square

root

filters. Mon.

Wea.

Rev., 142, 4499–4518,

676

Hemispheric Scale

Modes of Variability

NH

SH

GL

NH. Pacific

NH. Atlantic

Tropics

SH

PF-SIR

0.67

~0

0.24

0.23 (0.85)

0.67 (0.99)

0.25 (0.98)

~0 (0.95)

EAKF 5000 km

0.85

~0

0.61

0.57 (0.85)

0.78 (0.98)

0.43 (0.95)

0.25 (0.46)

EAKF 2000 km

0.74

~0

0.36

0.42 (0.67)

0.70 (0.99)

0.30 (0.97)

0.27 (0.23)

EAKF 500 km

0.51

~0

~0

~0 (0.95)

0.44 (0.99)

~0 (0.98)

0.30 (0.54)

677 678

Table 1. Correlation coefficients between different experiments and truth based on hemispheric scale averaged SST

679

time series from Figure 2; Correlation coefficients for PCs series between each experiment and the truth for the first

680

mode from Figure 7. Pattern correlation coefficients for the leading mode EOF between each experiment and the truth

681

(in brackets). The correlation coefficients that are not significant at 0.05 confidence level are marked as zero.

682

683 684

Figure 1. Observation locations and spatial reconstruction skill represented by RMSE and correlation coefficient for

685

4 experiments. a) shows the location of 47 proxies, the shading shows the seasonal natural variability/standard

686

deviation of the SST from the truth run. b) and c), d) and e), f) and g), h) and i) correspond to the four experiments:

687

PF-SIR, EAKF with localization 5000km, EAKF with localization 2000km, EAKF with localization 500km. b), d),

688

f), h) are RMSE divided by natural variability from a). c), e), g), i) are the spatial correlation map between each

689

experiment and the truth and the dotted area is not significant at 0.05 significance level.

690

691 692 693

Figure 2. Area averaged SST anomalies for southern hemisphere (g,h,i), northern hemisphere (d,e,f) and global (a,b,c);

694

Black curve is for truth; blue curve is for PF-SIR; red curve is for EAKF with localization scale 5000 km; orange

695

curve is for EAKF with localization scale 2000 km; green curve is for EAKF with localization scale 500 km. The

696

correlation coefficient with truth is shown for each experiment on a), d) and g); In the second column, the pink curves

697

are the 16 ensemble members from EAKF with localization scale 5000km; In the third column, the light blue curves

698

are for 48 ensemble members from PF-SIR.

699 700

Figure 3. Correlation coefficients between hemispheric averaged SST time series of EAKF- 5000 (red dot), EAKF-

701

2000 (magenta dot), EAKF-500 (green dot), PF-SIR (blue dot) and that of the truth at global scale (a), northern

702

hemisphere (b) and southern hemisphere (c). Each dot represents a 20-year, 40-year, or 100-year sample from the 200-

703

year-long simulation. The red and blue diamonds are the correlation coefficient of the 50-year-long experiments for

704

EAKF with 32 members and PF-SIR with 96 members.

705 706

Figure 4. First EOF field of North Pacific hemisphere (left column), North Atlantic (right column) ; The first row (a,

707

b ) is for the truth state, the second row (c, d) is for PF-SIR, the third row (e, f) is for EAKF with localization 5000km,

708

the forth row (g, h) is for EAKF with localization 2000km, and the last row (i, j) is for EAKF with localization 500km.

709

All figures are drawn in the same color scale and the unit is ℃. On top of each figure, the percentage of explained

710

variance is shown.

711

712 713

Figure 5. One-point correlation map calculated from seasonal mean model control at selected observation locations.

714

The black dot represents the observation location. The dotted area is not significant at 0.05 significance level.

715 716 717

Figure 6. First EOF field of the tropics ; a), b), c), d), and e) are for the actual state, PF-SIR, EAKF with localization

718

5000km, EAKF with localization 2000km, and EAKF with localization 500km respectively. All figures are drawn in

719

the same color scale and the unit is ℃.

720 721

Figure 7. Same as Figure 4. But for the southern hemisphere.

North Pacific

PF-SIR 0.23; EAKF: 0.57, 0.42, 0.04 2 0 -2 EAKF500

EAKF2000

truth

PF-SIR

EAKF5000

-4 0

50

0

50

North Atlantic

3

100

150

200

100

150

200

100

150

200

100

150

200

PF-SIR 0.67; EAKF: 0.78, 0.70, 0.44

2 1 0 -1 -2 -3 4

PF-SIR 0.25; EAKF: 0.43, 0.30, -0.11

Tropics

2 0 -2 -4 0

50

3

PF-SIR 0.02; EAKF: 0.26, 0.27, 0.30

2

SH

1 0 -1 -2 -3 0

50

Years

722 723

Figure 8. The normalized PCs for the first EOFs from Figure 3 - 5 for north Pacific, north Atlantic, tropics and

724

southern hemisphere. The black curve is for truth; the blue curve is for PF-SIR, the red curve is for EAKF with

725

localization scale 5000 km, the magenta curve is for EAKF with localization scale 2000 km, the green curve is for

726

EAKF with localization scale 500 km.

727 728

Figure 9. Spatial reconstruction skills for PF-SIR (left column) and EAKF-5000km (right column) on air surface

729

temperature (AST, a and b), 500hPa geopotential height (Z, c and d), and precipitation (PRECP, e and f).

730 731

Figure 10. Same as Figure 1 but for EAKF-5000km with 32 ensemble members (b and c), and PF-SIR with 96

732

ensemble members (d and e).

733 734

Figure 11. Same as Figure 2 a,d,g, but for EAKF-5000km with 32 ensemble members (red dashed line), and PF-SIR

735

with 96 ensemble members (blue dashed line).

736

737

Captions:

Hemispheric Scale

Modes of Variability

NH

SH

GL

NH. Pacific

NH. Atlantic

Tropics

SH

PF-SIR

0.67

~0

0.24

0.23 (0.85)

0.67 (0.99)

0.25 (0.98)

~0 (0.95)

EAKF 5000 km

0.85

~0

0.61

0.57 (0.85)

0.78 (0.98)

0.43 (0.95)

0.25 (0.46)

EAKF 2000 km

0.74

~0

0.36

0.42 (0.67)

0.70 (0.99)

0.30 (0.97)

0.27 (0.23)

EAKF 500 km

0.51

~0

~0

~0 (0.95)

0.44 (0.99)

~0 (0.98)

0.30 (0.54)

738 739

Table 1. Correlation coefficients between different experiments and truth based on hemispheric scale averaged SST

740

time series from Figure 2; Correlation coefficients for PCs series between each experiment and the truth for the first

741

mode from Figure 7. Pattern correlation coefficients for the leading mode EOF between each experiment and the truth

742

(in brackets). The correlation coefficients that are not significant at 0.05 confidence level are marked as zero.

743 744

Figure 1. Observation locations and spatial reconstruction skill represented by RMSE and correlation coefficient for

745

4 experiments. a) shows the location of 47 proxies, the shading shows the seasonal natural variability/standard

746

deviation of the SST from the truth run. b) and c), d) and e), f) and g), h) and i) correspond to the four experiments:

747

PF-SIR, EAKF with localization 5000km, EAKF with localization 2000km, EAKF with localization 500km. b), d),

748

f), h) are RMSE divided by natural variability from a). c), e), g), i) are the spatial correlation map between each

749

experiment and the truth and the dotted area is not significant at 0.05 significance level.

750 751

Figure 2. Area averaged SST anomalies for southern hemisphere (g,h,i), northern hemisphere (d,e,f) and global (a,b,c);

752

Black curve is for truth; blue curve is for PF-SIR; red curve is for EAKF with localization scale 5000 km; orange

753

curve is for EAKF with localization scale 2000 km; green curve is for EAKF with localization scale 500 km. The

754

correlation coefficient with truth is shown for each experiment on a), d) and g); In the second column, the light blue

755

curves are for 48 ensemble members from PF-SIR; In the third column, the pink curves are the 16 ensemble members

756

from EAKF with localization scale 5000km.

757 758

Figure 3. Correlation coefficients between hemispheric averaged SST time series of EAKF- 5000 (red dot), EAKF-

759

2000 (magenta dot), EAKF-500 (green dot), PF-SIR (blue dot) and that of the truth at global scale (a), northern

760

hemisphere (b) and southern hemisphere (c). Each dot represents a 20-year, 40-year, or 100-year sample from the 200-

761

year-long simulation. The red and blue diamonds are the correlation coefficient of the 50-year-long experiments for

762

EAKF with 32 members and PF-SIR with 96 members.

763 764

Figure 4. First EOF field of North Pacific hemisphere (left column), North Atlantic (right column) ; The first row (a,

765

b ) is for the truth state, the second row (c, d) is for PF-SIR, the third row (e, f) is for EAKF with localization 5000km,

766

the forth row (g, h) is for EAKF with localization 2000km, and the last row (i, j) is for EAKF with localization 500km.

767

All figures are drawn in the same color scale and the unit is ℃. On top of each figure, the percentage of explained

768

variance is shown.

769

Figure 5. One-point correlation map calculated from seasonal mean model control at selected observation locations.

770

The black dot represents the observation location. The dotted area is not significant at 0.05 significance level.

771 772

Figure 6. First EOF field of the tropics ; a), b), c), d), and e) are for the actual state, PF-SIR, EAKF with localization

773

5000km, EAKF with localization 2000km, and EAKF with localization 500km respectively. All figures are drawn in

774

the same color scale and the unit is ℃.

775 776

Figure 7. Same as Figure 4. But for the southern hemisphere.

777 778

Figure 8. The normalized PCs for the first EOFs from Figure 3 - 5 for north Pacific, north Atlantic, tropics and

779

southern hemisphere. The black curve is for truth; the blue curve is for PF-SIR, the red curve is for EAKF with

780

localization scale 5000 km, the magenta curve is for EAKF with localization scale 2000 km, the green curve is for

781

EAKF with localization scale 500 km.

782

783

Figure 9. Spatial reconstruction skills for PF-SIR (left column) and EAKF-5000km (right column) on air surface

784

temperature (AST, a and b), 500hPa geopotential height (Z, c and d), and precipitation (PRECP, e and f).

785 786

Figure 10. Same as Figure 1 but for EAKF-5000km with 32 ensemble members (b and c), and PF-SIR with 96

787

ensemble members (d and e).

788 789

Figure 11. Same as Figure 2 a,d,g, but for EAKF-5000km with 32 ensemble members (red dashed line), and PF-SIR

790

with 96 ensemble members (blue dashed line).

791 792

Suggest Documents