1 2 3 4 Typhoon-Position-Oriented Sensitivity Analysis. 5 Part I: Theory and Verification 6 7 8 9 10 11 12 13 14 15 Journal of the Atmospheric Sciences 16 (Accepted on Feb. 6, 2013). 17 18 19 20 Kosuke Ito 21 National University, Taipei, Taiwan and 22 Agency for Marine-Earth Science and Technology, Yokohama, Japan 23 24 Chun-Chieh Wu* 25 National Taiwan University, Taipei, Taiwan

26

27

28 *Corresponding author address: Chun-Chieh Wu, Department of 29 Atmospheric Sciences, National Taiwan University, No. 1, Sec. 4, 30 Roosevelt Rd., Taipei 106, Taiwan. 31 E-mail: [email protected]

1

32 Abstract

33 A new sensitivity analysis method is proposed for the ensemble

34 prediction system in which a (TC)-position is taken as a

35 metric. Sensitivity is defined as a slope of linear regression (or its

36 approximation) between state variable and a scalar representing the TC

37 position based on ensemble simulation. The experiment results illustrate

38 important regions for ensemble TC track forecast. The typhoon-position-

39 oriented sensitivity analysis (TyPOS) is applied to Typhoon Shanshan

40 (2006) for the verification time of up to 48 hours. The sensitivity field of

41 the TC central latitude with respect to the vorticity field obtained from

42 large-scale random initial perturbation is characterized by a horizontally

43 tilted pattern centered at the initial TC position. These sensitivity signals

44 are generally maximized in the middle troposphere and are far more

45 significant than those with respect to the divergence field. The results are

46 consistent with the sensitivity signals obtained from existing methods. The

47 verification experiments indicate that the signals from TyPOS

48 quantitatively reflect an ensemble-mean position change as a response to

49 the initial perturbation. Another experiment with Typhoon Dolphin (2008)

50 demonstrates the long-term analysis of forecast sensitivity up to 96 hours.

51 Several additional tests have also been carried out to investigate the

52 dependency among ensemble members, the impacts of using different

53 horizontal grid spacing, and the effectiveness of ensemble-Kalman-filter-

54 based perturbations.

2

55 1. Introduction

56 In terms of disaster prevention and mitigation associated with

57 tropical cyclones (TCs), the accuracy of track forecasts is critical because

58 forecasts of the intensity and the rainfall amount can result in large errors if

59 the track forecast is inadequate (Wu and Kuo 1999; Chan 2010). During

60 past decades, the physical processes governing the movement of TCs have

61 been intensively investigated (Wu and Emanuel 1993; Elsberry 1995; Chan

62 2010). The 48-h track error has been almost halved due to increased

63 observations and many improvements in numerical weather prediction

64 systems (National Hurricane Center, 2012). However, “busted” cases in

65 which track forecast errors exceed 1000 km in a 72-hour forecast still

66 occasionally occur (Yamaguchi et al. 2009, 2012; Japan Meteorological

67 Agency 2010). Therefore, reducing track forecast errors remains one of the

68 major issues for TC researchers and operational centers.

69 The errors in weather forecasts are mainly associated with two

70 factors: imperfections of the numerical model and errors in the initial

71 conditions (Yoden 2007). For a given model, additional observations from

72 critical regions are supposed to reduce uncertainties in the initial conditions

3

73 and lead to a better forecast. For the purpose of reducing errors in the track

74 forecast, aircraft-borne observations have been ‘targeted’ in the synoptic

75 environment of TCs. At present, operational surveillance observations of

76 TCs have been conducted by the National Oceanic and Atmospheric

77 Administration (NOAA) in the Atlantic basin since 1997 (Aberson and

78 Franklin 1999; Aberson 2003) and by the Dropwindsonde Observations for

79 Typhoon Surveillance near the Taiwan Region (DOTSTAR) project (Wu et

80 al. 2005) for selected western North Pacific TCs since 2003. In the summer

81 of 2008, the issue of targeted observation was further explored in the

82 Observing System Research and Predictability Experiment (THORPEX)

83 Pacific Asian Regional Campaign (T-PARC; Elsberry and Harr 2008;

84 Weissmann et al. 2011; Chou et al. 2011; Wu et al. 2012a,b; Huang et al.

85 2012).

86 Several techniques for targeted observation guidance have been

87 developed and compared to each other to identify “sensitive" regions

88 (Majumdar et al. 2006; Reynolds et al. 2007; Wu et al. 2009a). These

89 techniques include the total energy singular vector (SV) method used by

90 the Navy Operational Global Atmospheric Prediction System (NOGAPS;

4

91 Peng and Reynolds 2006; Chen et al. 2009), the European Centre for

92 Medium-Range Weather Forecasts (ECMWF; Buizza et al. 2007), the

93 Japan Meteorological Agency (JMA; Yamaguchi et al. 2009), and Yonsei

94 University (YSU; Kim and Jung 2009; Kim et al. 2011); the adjoint-derived

95 sensitivity steering vector (ADSSV) guidance (Wu et al. 2007, 2009b;

96 Chen et al. 2011); the ensemble transform Kalman filter (ETKF; Majumdar

97 et al. 2006, 2011); the analysis of ensemble deep-layer mean (DLM) wind

98 variance (Aberson 2003); and the Conditional Nonlinear Optimal

99 Perturbation (CNOP) method (Mu et al. 2009). While TC track forecasts

100 have been improved with additional observations in sensitive areas (e.g.,

101 Chou and Wu 2008; Yamaguchi et al. 2009; Chou et al. 2011), some

102 differences can be found among different methods (Majumdar et al. 2006;

103 Wu et al. 2009a). These differences stem from the different approaches to

104 detect the sensitivity on TC motion as well as the different numerical

105 models and configurations used (Wu et al. 2009b; Kim et al. 2011).

106 One potential way for further elaborating the sensitivity analysis

107 method is to employ TC position as a metric, whereas existing methods use

108 only TC steering flow (ADSSV) or total energy norms (SV, ETKF and

5

109 CNOP) at the verification time. The metrics employed in the existing

110 studies are presumably relevant to TC motion, and several studies have

111 shown that these methods can help capture the effect of synoptic features

112 and binary interactions (e.g., Peng et al. 2007; Wu et al. 2007).

113 Nevertheless, it is worth investigating a more direct evaluation of the

114 sensitivity field on TC tracks by using TC position as a metric.

115 In this study, we propose a typhoon-position-oriented sensitivity

116 analysis (TyPOS), which directly takes a TC position as a metric, for the

117 ensemble prediction system. This technique is based on the ensemble-based

118 sensitivity analysis method (Martin and Xue 2006; Ancell and Hakim 2007;

119 Torn and Hakim 2008, 2009; Liu et al. 2008; Gombos and Hansen 2008;

120 Sugiura 2010; Aonashi and Eito 2011). It has been applied to the sensitivity

121 analysis of the minimum sea level pressure (SLP) and the root mean square

122 error of SLP of TCs (Torn and Hakim, 2009) as well as the 1000-hPa

123 potential vorticity (Gombos et al. 2012). The ensemble-based sensitivity

124 analysis has also been used to correct the displacement error suitable for

125 real world observations using the data assimilation technique (Aonashi and

126 Eito, 2011). These approaches motivate us to calculate the sensitivity field

6

127 by employing TC position directly as a metric. Using this metric allows us

128 to bypass the requirement of defining the metric associated with TC motion,

129 and to perform sensitivity analysis without constructing tangent linear and

130 adjoint models.

131 In this study, we apply TyPOS to Typhoons Shanshan (2006) and

132 Dolphin (2008). The synoptic sensitivity field affecting the TC motion is

133 shown and verification tests are conducted to evaluate the response of the

134 TC displacement to initial perturbations. This work is the first attempt to

135 quantify the direction and distance of the TC vortex displacement directly

136 by using the sensitivity field. A companion paper (hereafter referred to as

137 Part II) will describe comparisons with other methods and provide physical

138 interpretations of the sensitivity field in detail.

139

140 2. Theoretical background

141 a. Review of ensemble-based sensitivity

142 TyPOS is based on the ensemble-based sensitivity analysis recently

143 developed by several researchers (Martin and Xue 2006; Ancell and Hakim

144 2007; Torn and Hakim 2008, 2009; Liu et al. 2008; Gombos and Hansen

7

145 2008; Sugiura 2010; Aonashi and Eito 2011). Here, we briefly outline this

146 sensitivity analysis technique in a generalized manner.

147 Let x and j be an n degree-of-freedom input vector and output

148 scalar, respectively. The input vector represents the perturbations in the

149 physical variables at the initial time and the output scalar is a forecast

150 metric. Considering m realizations of the input vector,

151 xi(x1,i,x2,i,...,xn,i), where i(i1,2,...,m) represents the index of each

152 ensemble member, with corresponding realizations of the output scalar,

153 ji , we define matrix X and vector j as follows: 1 T 154 X x1,x2,...,xm, (1) m1 1 T 155 j j1,j2,...,jm. (2) m1

156 Matrix is composed of xi and the vector is composed of ji

157 where  indicates deviations from ensemble mean values, that is,

158 xi xi x and ji ji  j . The angle bracket  denotes the

159 ensemble mean and superscript T denotes the transpose. Note that the

160 deviations are not necessarily infinitesimal, while Ancell and Hakim (2007)

161 introduced the assumptionJ/xJ/x.

162 The generalized ensemble-based sensitivity is defined as follows

8

163 (Gombos et al., 2008):

j  164 λx X j , (3)

165 where the superscript cross ( ) indicates the Moore-Penrose generalized

166 inverse matrix1, which can be computed from the equation below.

167 X  VDUT , (4)

168 where X UDVT is the singular vector decomposition of X . The

169 generalized inverse matrix D is formed by replacing every nonzero

170 diagonal entry of D by its reciprocal and transpose the resulting matrix.

j 171 In this paper, the vector λ x is referred to as “sensitivity” and the k-

λ j (k1,2,...,n) 172 th component of is denoted as xk . The reason why

173 is called the “sensitivity” is that it represents the generalized solution of

174 inverse problem associated with a first-order relationship between j and

175 X ,

176 j Xλˆ, (5)

177 which means the vector λˆ multiplied by initial perturbations in physical

1 A Moore-Penrose generalized inverse matrix of X is defined as a matrix satisfying all of the

       following four criteria: XX X  X, X XX  X, (XX )*  XX and (X X)*X X, where the asterisk

represents the adjoint matrix. The matrix exists and is unique. See more details in Menke (1989).

9

178 variables equals to perturbations in the forecast metric at the verification

179 time.

180 If the number of ensemble members exceeds the degrees-of-freedom,

181 i.e., m  n , a purely over-determined problem and the inverse of (XTX)-1

ˆ j 182 exists, the vector λ  λ x is the slope of regression line obtained from

183 minimizing deviations as shown in Fig. 1 (Menke, 1989). In other words,

λ j j x 184 xk represents the ratio of changes in to the change in k based on

j 185 ensemble simulations. The vector λ x is the same as the adjoint sensitivity

186 if the time evolution of initial perturbation is completely equivalent to that

187 described by the tangent linear model and is a continuous function of

188 x (Ancell and Hakim, 2007).

189 If the degrees-of-freedom of exceeds the number of ensemble

190 members, i.e., m  n , a purely under-determined problem, the vector

ˆ j 191 λ  λ x represents the solution which satisfies Eq. (5) and yields the

ˆ L 192 minimum of λ in the 2 norm (Menke 1989). This type of solution is

193 termed as a minimum length solution and is frequently used in tackling

194 inverse problems. Although the vector can still be interpreted as a

195 slope with respect to satisfying j  Xλˆ , the additional condition

10

ˆ 196 (minimization of λ ) is introduced to fully determine all the n-elements of

197 the sensitivity field from m realizations. Thus, we do not expect that this

198 sensitivity would appropriately reflect the impact of changes in the input

199 vector, x , on changes in the metric, j , when the number of ensemble

200 members, m , is far smaller than the degrees-of-freedom of , n .

201 If the inverse matrix of XTX is known, the sensitivity can be

202 obtained without executing the singular value decomposition in equation

203 (4). By multiplying both sides of equation (5) by XT, we obtain

T Tˆ 204 XjXXλ. (6)

205 Since an element in the a -th row and the b-th column of XTX is 1 m 206 xa,ixb,i , can be replaced by the ensemble covariance matrix. m1i1

207 In particular, when the ensemble covariance matrix is diagonal, the

208 equations (3) and (4) are reduced to the simplified equation below:

1 m xk,iji 209 λj m1i1 , (7) xk 2 k

2 2 210 where  k represents ensemble variance .

211

2 Note that Eq. (7) is used as a definition of ensemble-based sensitivity for both diagonal and off-diagonal matrix in Ancell and Hakim (2007).

11

212 b. TyPOS

213 TyPOS is an application of the ensemble-based sensitivity analysis

214 designed for the ensemble forecast system. The input vector x represents

215 perturbations in the physical variables at the initial time and the output

j 216 scalar is a scalar representing the vortex position such as longitude and

217 latitude of the TC center at the verification time. TyPOS does not assume

218 that the perturbations are infinitesimal. It fits our objective of employing

219 TC position as a metric because it is not a differentiable function when

220 reproduced in a dynamical model.

221 One problem is that the ensemble-based sensitivity is not expected to

222 be a good indicator for sensitivity when the number of ensemble members

223 is far smaller than the degrees-of-freedom of . If we consider the

224 sophisticated three-dimensional numerical model with degrees-of-freedom

225 n', it is not practical to perform simulations as much as times.

226 To resolve this problem, we consider a reduced-dimension space in

227 which the n-elements ( n  n') of used for sensitivity analysis are

228 sampled with larger grid spacing compared with that of the original

229 numerical model in this work. The horizontal grid spacing can be greater

12

230 than a few hundred kilometers if the aim is dealing with the synoptic

231 features. In this work, the grid spacing used to define the sensitivity field

232 differs from the grid spacing used to define the forecast metric. The grid

233 points on the coarse mesh used for sensitivity calculation are referred to as

234 the reduced grid points, while the grid points on the original model mesh

235 used to perform the time integration are referred to as the model grid points.

236 Here, initial perturbations in the reduced grid points are interpolated into

237 the model grid points by cubic spline. The use of a basis function

238 corresponding to the fast developing modes is another option aside from

239 employing the large-scale perturbations, which will be discussed in section

240 6b.

241 TyPOS has the potential to quantify the displacement distance and

242 direction of ensemble-mean TC position. Suppose that the sensitivity

243 approximates the ratio of changes in j to the change in xk , ensemble-

244 mean change in the output metric as a response to the initial perturbation is

245 evaluated as follows ( denotes the perturbation relative to the control

246 run):

1~j 247 jx'x', (8) S

13

248 where  denotes the summation over all the model grid points, and x'

249 represents the changes in the initial state defined at the model grid points.

250 In order to complete the manipulation between the initial perturbations

251 defined on model grid points and the sensitivity defined on reduced grid

252 points, we introduce the interpolated sensitivity field defined on the model

~j 253 grid points, x' , and the scaling factor S. The scaling factor is defined as

~ 254 Sλ λ (in L1 norm), which is based on the assumption that the influence

255 of the initial changes in a reduced grid point is equivalent to the integrated

256 influence of uniform initial condition changes in corresponding model grid

257 points, i.e., it has a large value when the reduced grid points are coarsely

258 sampled from the model grid points. Eq. (8) is useful for evaluating the

259 magnitude of displacement as a response to changes in the initial state

260 variables and for determining the priority of the targeted observations,

261 although the expected displacement distance should be regarded as an

262 approximation. This is due to the following reasons: the truncation error

263 inherent in the projection, dependency on perturbation magnitudes, and the

264 assumption that impacts are calculated by simple linear superposition of

265 impacts from single grid points.

14

266 As stated in Torn and Hakim (2008), the reliability of sensitivity

267 signals in TyPOS can be measured using a t-test. The t-test is relevant to the

268 rejection of the null hypothesis that the slope of the output metric, j , with

269 respect to the single variable, xk , is zero (Wilks 2005). A higher

270 confidence level is obtained with the increasing number of ensemble

271 members or with statistically smaller deviations from the regression line if

272 the values of the slope are equal. It is critical to distinguish meaningful

273 signals from statistical noise, in particular, when the number of ensemble

274 simulations is far smaller than the degrees-of-freedom of initial

275 perturbations as demonstrated in section 6a.

276

277 c. Summary of TyPOS

278 Before applying the proposed method to a real-case calculation, the

279 features and procedures of TyPOS are summarized here. In this

280 methodology, TC position is taken as a forecast metric, and thus it is a

281 direct method for a given ensemble prediction system. In addition to the

282 clear objectivity, this method has several advantages. The method provides

283 the possibility to evaluate the ensemble-mean vortex displacement

15

284 direction and distance without using the tangent linear and adjoint models.

285 To alleviate the burden posed by the enormous computational resources

286 needed for obtaining the significant signals, we consider the reduced-

287 dimension space and, if needed, to distinguish meaningful signals from

288 noise using t-tests.

289 The procedure of TyPOS is as follows: (i) Generating m realizations

290 with n degrees-of-freedom in the reduced grid points and projecting them

291 onto the model variable space by cubic spline interpolation, (ii) performing

292 the ensemble simulations m times after adding interpolated perturbation

293 to the reference initial state, (iii) calculating the sensitivity by using the m

294 realizations of the resultant metric that represents the TC position [Eqs. (3)

295 and (4), or Eq. (6)], (iv) (optional) evaluating the displacement of the

296 vortex position by multiplying the changes in the initial state variable with

297 the sensitivity field after the interpolation into the model grid points [Eq.

298 (8)], and (v) (optional) conducting t-tests to distinguish reliable signals

299 from noise.

300 TyPOS can directly reflect the sensitivity field of the metric relevant

301 to TC position for a given model and initial perturbations if the number of

16

302 ensemble members is sufficiently large. However, TyPOS is still subject to

303 a variety of uncertainties in terms of specifying initial perturbations. In this

304 study, limitations arise from the fact that dynamically balanced large-scale

305 perturbations are employed as initial perturbations. A fine scale feature is

306 briefly discussed in section 6, although detailed examination is beyond the

307 scope of this work.

308 Note that TyPOS is based on the linear regression obtained from

309 ensemble simulation and the perturbation is not necessarily infinitesimal.

310 Therefore, the validity of TyPOS is not limited to the time scale of tangent

311 linear assumption as in any singular vector- or adjoint-based method.

312 Nevertheless, care should be taken in interpreting the sensitivity, and the

313 influence of nonlinear growth of perturbations requires further study.

314

315 3. Experimental Design

316 a. Model Configuration and time schedule

317 In order to assess the feasibility of TyPOS, we apply this method to

318 Typhoon Shanshan, for which sensitivity analyses based on different

319 targeted methods have already been conducted (Wu et al. 2009a,b;

17

320 Reynolds et al. 2009; Chen et al. 2011) and Typhoon Dolphin, for which

321 the spread of ensemble forecast TC track was reported to be large (from

322 JMA 2009). The verification time is set to 12, 24 and 48 hours for

323 Shanshan in order to compare with the existing studies, and to 24, 48 and

324 96 hours for Dolphin in order to investigate its applicability to long-term

325 sensitivity analyses with large uncertainties in track forecasts.

326 In this study, the Advanced Research version of Weather Research

327 and Forecast (ARW-WRF) model version 2.2.1 is employed to conduct the

328 numerical simulation. The model configuration is the same as the doubly-

329 nested domain calculations of Wu et al. (2010). The coarse domain

330 (Domain 1) has 150x120 grid points (longitude by latitude) with the

o o 331 horizontal grid spacing of 54 km centered at 120.0 E and 27.5 N, while the

332 fine-meshed domain (Domain 2) has 120 x 150 grid points with the

o o 333 horizontal grid spacing of 18 km centered at 124.9 E and 27.1 N. The

334 model contains 35 vertical levels in the terrain following sigma coordinate.

335 The domains and TC position from the JMA best track are shown in Fig. 2.

336 The prognostic state variables are perturbation of potential

337 temperature, geopotential and dry air mass in a column from the basic state

18

338 as well as horizontal and vertical wind fields. Six mixing ratios (water

339 vapor, cloud water, cloud ice, rain, snow and graupel) are also

340 prognostically calculated based on the WRF single moment six-class

341 graupel microphysics scheme (Hong et al. 2004; Hong and Lim 2006).

342 Other parameterization schemes include the rapid radiative transfer model

343 scheme (Mlawer et al. 1997) for longwave radiation, the simple shortwave

344 scheme Dudhia (1989) for shortwave radiation, and the YSU planetary

345 boundary layer scheme (Hong et al. 2006). The Grell-Devenyi ensemble

346 scheme (Grell and Devenyi 2002) is used for the parameterization of

347 cumulus convection in both Domains 1 and 2.

348 To reproduce a realistic TC vortex, we performed a one-week data

349 assimilation with the EnKF technique prior to the sensitivity analysis. This

350 data assimilation procedure starts from the field of National Centers for

351 Environmental Prediction (NCEP) final analysis. The WRF-based EnKF

352 data assimilation system used in this study is similar to the system in Meng

353 and Zhang (2007, 2008a,b) and has been developed to adjust TC central

354 position and pressure in order to fit the best track (Wu et al., 2010). In this

355 study, the number of ensemble members in the EnKF system is 100. We

19

356 regard 0000 UTC 15 September 2006 (1200 UTC 11 December 2008) as

357 the initial time of the sensitivity analysis for Shanshan (Dolphin) and

358 regard ensemble-mean state of the EnKF analysis field at this moment as

359 the initial reference state. For simplicity, we assume that this initial time is

360 the same as the observation time (i.e., no lead-time forecast experiment).

361 For the presentation of results, the number of hours beginning from the

362 initial time of the sensitivity analysis is used. The time schedule of the

363 experiment is summarized in Fig. 3. In fact, we calculate the sensitivity

364 both of latitude and longitude at the model grid point where the minimum

365 pressure is achieved at the verification time. However, for concise

366 presentation, we will present the only sensitivity of the latitude (longitude)

367 for Shanshan (Dolphin) unless otherwise noted.

368

369 b. Perturbations for sensitivity analysis

370 In order to construct the sensitivity field, u and v at the initial

371 time are perturbed. As a first step, the magnitude is simply determined so

-1 - 372 that the standard deviation becomes an uniform value of 1.8 m s (1.4 m s

1 373 ) for Shanshan (Dolphin), which is the same as the area-averaged spread of

20

374 the 100 EnKF members in the middle troposphere at the initial time

375 (figures not shown).

376 In this study, initial perturbations are generated independently in

377 each reduced grid point and thus t-tests are applied independently in each

378 reduced grid point except in section 6b. The benefit of using the

379 uncorrelated perturbations is that we can calculate the sensitivity based on

380 the simplified equation (7). Independent Gaussian random numbers in each

381 reduced grid point are generated by using Mersenne Twister method and

382 Box-Mullar Transform (Box and Muller 1958; Matsumoto and Nishimura

383 1998) and are added to the initial reference state of these variables. It is

384 used to construct the sensitivity field with respect to u and v. Although the

385 relative vorticity  and divergence  are not prognostic state variables

386 in this model, these are essential variables for clear physical interpretation.

387 Thus, the sensitivity field with respect to u and v is converted to one with

388 respect to relative vorticity and divergence as in Kleist and Morgan

389 (2005) and Wu et al. (2007).

390 The horizontal reduced-grid spacing is set to 540 km regardless of

391 the model grid spacing. This coarse grid spacing used in this study is

21

392 presumably sufficient in resolving synoptic features of a few thousand

393 kilometers, while not reflecting the fine-scale structures of the sensitivity

394 field. The vertical grid points are sampled on  = 0.99885, 0.8445, 0.5060,

395 0.2879, 0.101. The potential temperature and geopotential fields are also

396 changed so that the perturbation fields are close to geostrophic and

397 hydrostatic balances as a response to initial perturbations of vorticity. For

398 instance, the standard deviation of perturbations in potential temperature

2 -2 399 (geopotential) is 0.12 K (30.1 m s ) at 20N and = 0.8445, and 0.06 K

2 -2 400 (14.2 m s ) at 20N and = 0.5060 for Shanshan. In order to obtain

401 temperature and geopotential fields, we follow the same procedure as in

402 Komaromi et al. (2011). We first calculate the perturbation in stream

403 function by solving Poisson’s equation using a relaxation method and then

404 calculate the temperature and geopotential perturbation field consistent

405 with the stream function.

406 The number of ensemble simulations ( m ) is 800. However, 170

407 simulations for Dolphin are not used for sensitivity analysis because the

408 maximum wind speed of the cyclone is reduced below the tropical storm

-1 409 intensity of 17.2 m s during their lifetime. This threshold is introduced to

22

410 strictly avoid false detection of TCs. The degrees-of-freedom of the initial

411 perturbations (n) is 1800 for Shanshan (15x12x5 reduced grid points in the

412 zonal, meridional and vertical direction for the two components, u and

413 v ) and 3600 for Dolphin (24x15x5x2). Therefore, the problems are

414 formulated as underdetermined ones.

415

416 4. Sensitivity analysis for Shanshan

417 a. Synopsis and simulated track

418 Figure 4 shows the center position of Shanshan based on the JMA

419 best track. It formed to the west of Guam and moved northwestward and

420 started to recurve on 13 September, turning northward from its original

421 westward path. The TC later began to accelerate northeastward on 16

422 September. During the period of the current numerical experiment, there

423 are several synoptic features around the TC region. Figures 5a,d show that

424 the location of Shanshan was between the midlatitude trough over

425 northern-central China and the anti-cyclonic flow at the initial time (0000

426 UTC 15 September). On 16 September, Shanshan moved swiftly as it

427 started to interact with a midlatitude trough over eastern China (Figs. 5b,e).

23

428 On 17 September, Shanshan has become embedded in the midlatitude

429 trough in the 500-hPa fields (Figs. 5c,f).

430 The simulated tracks of ensemble members from 0000 UTC 15

431 September to 0000 UTC 17 September are overlaid with TC symbols in Fig.

432 4. This shows that the TC central position is reproduced with a left-of-track

433 bias. Ensemble spreads of the forecasted TC central longitude and latitude

434 at 0000 UTC 17 September are 0.422 and 0.384 degree (a distance of 39

435 km and 43 km), respectively.

436

437 b. Ensemble-based sensitivity

438 Figure 6 shows the sensitivity field of the TC central latitude with

439 respect to the vorticity field at three vertical levels ( = 0.8445, 0.5060,

440 0.2879) for the verification time of 12, 24 and 48 hours. The values are

441 horizontally interpolated into the model grid points with a cubic spline. The

442 circles superposed in these figures indicate the t-test results. The sensitivity

443 related to the vorticity field is less significant at = 0.99885 and =

444 0.101 than at = 0.8445, 0.5060, 0.2879 (figures not shown). The signals

445 are the strongest at the verification time of 48 h, which reflects the fact that

24

446 the signals are relevant to the ratio of the change in the TC central latitude

447 at the verification time as a response to the unit initial change. The

448 evolution of the initial perturbation brings about large ensemble-mean

449 changes in the TC central latitude, particularly associated with the swift

450 passage as the TC interacts with midlatitude troughs. Another notable

451 feature is that the major signals are sufficiently far away from the lateral

452 boundary, implying that the regional model is able to reasonably reproduce

453 the TC track, at least within this time scale.

454 Here, we briefly outline the structure of the sensitivity field. Further

455 investigations on the physical interpretation and comparison of other

456 methods will be presented in Part II. Figure 6 shows that the largest

457 sensitivity near the TC center appears at  = 0.5060, with values

458 generally larger than those at = 0.2879. These are consistent with

459 results from previous sensitivity studies for Shanshan obtained with the

460 ADSSV and SV methods (Wu et al. 2009b; Reynolds et al. 2009), implying

461 that TC motion is more sensitive to changes in the middle troposphere.

462 For the sensitivity field at the verification time of 12 h, the north-

463 south dipole pattern appears to be centered at the initial TC position. A

25

464 vertically coherent structure is present from the lower troposphere to the

465 middle troposphere with the largest sensitivity occurring in the latter. This

466 dipole pattern is to some extent relevant to the latitudinal displacement of

467 the vortex. This implies that the appropriate positioning of the vortex is an

468 important element for improving short-term scale track predictions (Ito et

469 al. 2013). It is notable that the sensitivity signal of meridional velocity in

470 ADSSV (Wu et al. 2007)3 at the verification time of 12 h exhibits a west-

471 east dipole pattern in contrast to the signals in TyPOS. Physical

472 interpretation of this difference will also be investigated in Part II.

473 The sensitivity field exhibits a horizontally tilted pattern centered at

474 the initial TC position with respect to the vorticity field in the middle

475 troposphere. This pattern is typically found in ADSSV-based sensitivity

476 analyses (Wu et al. 2007, 2009b; Chen et al. 2011). It is intriguing that the

477 negative sensitivity signals are well developed from east to southwest

478 relative to the initial TC position with extended verification time, which is

479 consistent with the stronger sensitivity signals for Shanshan obtained from

480 ADSSV (Wu et al. 2009b; Chen et al. 2011). The confidence level is above

3 To avoid misunderstanding, note that the directions of the vectors in TyPOS and ADSSV have different physical implications.

26

481 99.9%. This feature is also detected in the SV-based sensitivity analysis at

482 the recurvature stage for Shanshan (Reynolds et al. 2009).

483 Besides the horizontally tilted pattern in the middle troposphere, the

484 sensitivity field is elongated northwestward in the upper layer at the

485 verification time of 48 h, presumably due to the strong southeastward wind.

486 This feature is also detected and investigated as a sign of interactions

487 between the TC and midlatitude trough from sensitivity signals of ADSSV

488 (Wu et al. 2009b) and SV (Reynolds et al. 2009). Under close inspection,

489 our signal around the midlatitude trough is not distinct in the middle

490 troposphere in contrast to the ADSSV- and moist SV-based sensitivity

491 signals.

492 The confidence levels of the sensitivity signals with respect to the

493 divergence field are mostly below 99% in the lower, middle and upper

494 troposphere (figures not shown). These sensitivity signals are not organized

495 in synoptic scale, suggesting that the vorticity field is more relevant to the

496 TC motion at least in comparison to the divergence field (consistent with

497 results in Wu et al. 2007). As mentioned above, the sensitivity signals

498 obtained from the existing SV- and ADSSV-based methods are fairly

27

499 consistent with the signals in TyPOS, although some differences are also

500 identified.

501

502 c. Verification experiment

503 To ensure that the sensitivity signals obtained from TyPOS are

504 relevant to the motion of Shanshan, additional ensemble experiments are

505 conducted. One hundred members are randomly sampled from eight

506 hundred members used for the sensitivity calculation. This set of 100

507 ensemble members is referred to as CTRL. We perform ensemble

508 experiments in which three patterns of positive vorticity are added to each

509 initial state of CTRL in the prescribed region. To conduct an ensemble run,

510 the positive vorticity perturbations (without the divergence) are

511 transformed into horizontal velocity perturbations (u,v), which are the

512 model prognostic variables, via calculations of the stream function by

513 solving Poisson’s equation. The amplitude of the initial perturbation of u

514 and v is adjusted to the size of standard deviation consistent with the

515 sensitivity analysis.

516 Here, we investigate the validity of sensitivity field in the middle

28

517 troposphere with the verification time of 48 h. Two experiments are

518 associated with the strongest sensitivity signals around the initial TC

519 positions and the other experiment is associated with the midlatitude trough

520 in the middle troposphere where differences arise between TyPOS and

521 existing methods. We refer to these three experiments as follows: (i)

522 WTILT is the experiment in which the positive vorticity perturbation is

523 added to the regions consistent with the horizontally tilted pattern west to

524 the initial TC position from  = 0.66175 to = 0.4082, (ii) ETILT is the

525 same as WTILT but with the horizontally tilted pattern east to the initial TC

526 position, (iii) TROUGH is the same as WTILT and ETILT but with the

527 perturbation near the midlatitude trough. Figures 7a-c indicate the initial

528 perturbations of additional horizontal wind in ETILT, WTILT and

529 TROUGH, respectively. Based on the definition of sensitivity and Fig. 6f,

530 we can expect an ensemble-mean northward (southward) displacement

531 relative to CTRL in the WTILT (ETILT) experiment and a smaller

532 displacement in the TROUGH experiment.

533 According to equation (8), the expected ensemble-mean changes

534 (Δ and Δ) are evaluated as a response to the additional

29

535 horizontal wind perturbation at the initial time. As mentioned in section 2b,

536 we multiply the magnitude of perturbations by the sensitivity both of

537 latitude and longitude which are interpolated to the model grid points using

538 a cubic spline.

1~LON~LON1~LON~LON LONuvuv 539 Su vSu v (9) D1D1 D2D2

1~LAT~LAT1~LAT~LAT LATuvuv 540 Su vSu v (10) D1D1 D2D2

541 where the subscripts D1 and D2 represent domain 1 and domain 2, and

S S 542 D1 and D2 are set to constant values of 617.7 and 6273.5, respectively.

543

544 d. Results of the verification experiment

545 Figure 8 shows a typical example of time evolution of perturbation in

546 a member of the ETILT experiment (ETILT_m001) relative to a

547 corresponding member of the control experiment (CTRL_m001).

548 Perturbations are developed in a broader region within the first 12 hours

549 (Fig. 8b). They continue to grow around the TC with maximum values of

-1 550 about 20 m s (Figs. 8c,d), appearing to exhibit a cyclonic circulation

551 southwest to the TC position and an anti-cyclonic circulation northeast at

30

552 the verification time. This is consistent with the southwestward

553 displacement of the TC in ETILT_m001 relative to that in CTRL_m001.

554 Figure 9a shows the ensemble-mean TC positions in different

555 experiments. The members in the WTILT run tend to move northward

556 relative to the CTRL run, while the members in the ETILT run tend to

557 move southward. This finding is consistent with the results of the

558 sensitivity analysis. Ensemble-mean displacement in the TROUGH

559 experiment remains small as expected. It does not necessarily mean that the

560 existing methods overestimate the sensitivity associated with the

561 midlatitude trough in the middle troposphere because the sensitivity is also

562 dependent on the numerical model and the characteristics of initial

563 perturbations. Nevertheless, the results demonstrate that TyPOS accurately

564 reflects the sensitivity of TC position as a response to large-scale initial

565 perturbation for a given ensemble forecast system.

566 Figure 9b shows the frequency distributions of TC central latitude of

567 the 100 members among different ensemble experiments. It is obvious that

568 the ensemble-mean TC position in WTILT (ETILT) is located to the north

569 (south) of the TC position in CTRL, while the distribution overlaps among

31

570 different experiments. The current method is shown to be relevant to the

571 changes in probability density function of TC positions.

572 Figure 9c illustrates the ensemble-mean TC position change relative

573 to CTRL with nonlinear model calculation. The TC displacement estimated

574 from TyPOS quantitatively agrees well with the ensemble-mean

575 displacements in the perturbed experiments (Fig. 9d). We calculate the

576 amplification factor defined as the ratio of total kinetic energy (TKE) of

577 ensemble-mean perturbation integrated over a 1000 km by 1000 km square

578 centered at the TC position at the verification time to that over the entire

579 domain at the initial time. The amplification factors are 1.47 for WTILT,

580 3.73 for ETILT and 0.13 for TROUGH at t = 48 h. It is consistent with the

581 largest development of perturbations near Shanshan in the ETILT

582 experiment.

583 Overall, these features indicate the usefulness of TyPOS. The method

584 can be used to evaluate the expected direction and distance of the vortex

585 displacement as a response to the changes in the initial state and to identify

586 regions that have greater impacts on an ensemble of realizations of the TC

587 forecast position under the probabilistic framework.

32

588

589 5. Sensitivity analysis for Dolphin

590 a. Synopsis and the simulated track

591 Figure 10 shows the center position of Dolphin in the JMA best track.

592 Dolphin turned into a tropical storm west of the Mariana Islands at 0600

593 UTC on 11 December 2008. After turning sharply to the north around 0000

594 UTC on 15 December, it began to move northeastward on 16 December.

595 The simulated tracks of ensemble members from 1200 UTC 11 December

596 to 1200 UTC 15 December are overlaid with TC symbols in Fig. 10.

597 Simulated tracks generally exhibit early recurvature to the north, while TC

598 position varies greatly at the forecast time of 96 h. Ensemble spreads of the

599 forecasted TC central longitude and latitude at 1200 UTC on 15 December

600 are 1.63 and 1.33 degree (a distance of 172 km and 148 km), respectively.

601 Figure 11 shows that Dolphin was embedded in the easterly wind and

602 was located west of the anti-cyclonic flow at 1200 UTC on 11 December

603 (initial time of the sensitivity analysis). The anti-cyclonic flow and the

604 midlatitude trough intensified around 14 December and the simulated TCs

605 swiftly moved northeastward toward the direction of flow and the trough.

33

606

607 b. Ensemble-based sensitivity

608 Figure 12 shows the sensitivity field of the TC central longitude with

609 respect to the vorticity field at three vertical levels ( = 0.8445, 0.5060,

610 0.2879) for the verification time of 24, 48 and 96 h. The values of

611 sensitivity field at the verification time of 24 and 48 h are large as

612 compared to the sensitivity of the forecasted latitude of Shanshan. This

613 suggests that ensemble-mean TC position change is more sensitive to the

614 initial perturbation. The sensitivity with respect to the vorticity field is less

615 significant at = 0.99885 and = 0.101 and the same tendency is

616 observed in the sensitivity with respect to the divergence field as described

617 in section 4 (figures not shown).

618 The largest sensitivity appears at = 0.5060. The values at =

619 0.8445 and = 0.2879 are generally smaller than those at = 0.5060

620 regardless of the verification time. While the confidence level is generally

621 low compared to the case of Shanshan, some sensitivity signals have a

622 confidence level over 90% at the verification time of 96 h.

623 For the sensitivity field at the verification time of 24 h, the east-west

34

624 dipole pattern appears to be centered at the initial TC position. The positive

625 sensitivity area extends from the east of the initial position to southern

626 Taiwan in the middle troposphere at the verification time of 48 and 96 h.

o o 627 The significant signals are found around 120-130 E and 15-25 N, located at

628 the southern part of the westerly jet, which illustrates the influences of the

629 jet on the TC track.

630

631 c. Verification experiment

632 Four verification experiments are conducted to ensure that the

633 sensitivity signals in the middle troposphere are relevant to the TC position

634 at the verification time of 96 h. Referred to as P1, P2, N1 and N2, the

635 experiments correspond to the significant sensitivity signals around the TC.

636 The procedure of the four experiments is the same as that in section 4c, i.e.,

637 the positive vorticity perturbations (without divergence) from  =

638 0.66175 to = 0.4082 are added to the initial field of the 100-member

639 ensemble run. The initial perturbations of the horizontal wind are shown in

640 Figures 13a-d. Based on the definition of sensitivity and Fig. 12f, we can

641 expect to observe an ensemble-mean eastward (westward) displacement

35

642 relative to CTRL in P1 and P2 (N1 and N2) experiments.

643 The expected ensemble-mean changes (LON and  LAT) as a

644 response to the constant perturbation in the initial state are evaluated

645 according to the equation below:

1~LON~LON LONuv 646 Su v  (11) DD

1~LAT~LAT LATuv 647 Su v  (12) DD

S 648 where the subscript D represents the domain, and D is set to a constant

649 value of 630.

650

651 d. Results of the verification experiment

652 Figures 14a,b show ensemble-mean tracks of each experiment. It is

653 clear that ensemble-mean displacement at the verification time is consistent

654 with TyPOS. TC tends to move eastward in P1 and P2 experiments relative

655 to the CTRL, while the ones in the N1 and N2 run tend to move westward.

656 These tendencies are connected to earlier (later) recurvature in P1 and P2

657 (N1 and N2) experiments. Figures 14c,d show the frequency distributions

658 of longitude among different ensemble experiments. Although the

36

659 frequency distributions are heavily overlapped, the ensemble-mean value is

660 different among the different experiments.

661 Figure 14e illustrates the ensemble-mean TC position displacement

662 calculated in a nonlinear model. It is qualitatively consistent with the

663 evaluation by using TyPOS (Fig. 14f), with a significant signal only

664 slightly above 80%, indicating a difference from the case of Shanshan

665 which yields a rather quantitative result. Nevertheless, these results show

666 potential of TyPOS for long-term forecast sensitivity analyses. The

667 amplification factors with the verification time of t=96 h are 3.78 for N1,

668 1.01 for N2, 5.06 for P1 and 0.79 for P2. These figures correspond to the

669 large displacement in N1 and P1.

670

671 6. Discussion

672 a. Dependency on ensemble size and resolution

673 It is important to investigate the dependency of the sensitivity

674 analysis on the number of ensemble members. The sensitivity analysis is

675 conducted based on 400, 200 and 100 ensemble members for the same case

676 (Typhoon Shanshan) described in sections 3 and 4. Figure 15 shows the

37

677 sensitivity field at  = 0.5060 with respect to the vorticity field,

678 superposed with the results of significance t-test. The north-south dipole

679 pattern at the verification time of 12 h and the horizontally tilted patterns

680 centered at the initial TC position are found in the sensitivity field with

681 fewer number of ensemble members.

682 However, some spurious features appear away from the TC center

683 and the confidence level of the signals decreases with a smaller number of

684 ensemble members. One encouraging result is that noise caused by the

685 decreasing number of ensemble members is not detected as significant in

o o 686 most of the cases, although some signals (e.g., 92 E and 32 N in Fig. 15l)

687 indicate a 90% confidence level.

688 So far, the reduced grid spacing is set to 540km, which is too coarse

689 to investigate fine scale structures. Another sensitivity analysis is

690 performed by halving the horizontal reduced-grid spacing to 270 km for

691 Typhoon Shanshan (2006). With the exception of the reduced-grid spacing,

692 the settings are the same as in the original one described in sections 3 and 4.

693 Figure 16 shows the sensitivity fields with respect to the vorticity fields.

694 The sensitivity fields are consistent with the findings shown in section 4 in

38

695 that the horizontally tilted pattern is found near the initial TC position.

696 However, the sensitivity field is contaminated by statistical noise and the

697 signals are seldom considered significant.

698 This suggests that TyPOS starting from large-scale perturbations can

699 serve as a promising method for both targeted observation operations and

700 research purposes, given the large number of ensemble members. In TyPOS,

701 an additional assumption is introduced to fully determine n-elements of the

702 sensitivity field from m realizations as explained in section 2a. Thus, the

703 sensitivity field is not likely to reflect the relationship between input

704 physical variables and the output scalar as the number of degrees-of-

705 freedom increases. Recent developments in large parallel computer systems

706 with thousands of processors make it possible to obtain sensitivity fields

707 from a very large number of nonlinear forward model runs with full physics

708 (Martin and Xue 2006). However, without enough computational resources,

709 this method provides less reliable features or is merely useful for capturing

710 some features in the further reduced-dimension space. In terms of the

711 operational ensemble forecast, it would be possible to use a basis function

712 corresponding to the fast developing mode as discussed below or to shrink

39

713 the domain size of initial perturbations along with the reduction of

714 ensemble size.

715

716 b. The use of EnKF-based perturbations

717 So far, large-scale random initial perturbations are used to illustrate

718 the feasibility of TyPOS. However, the use of faster developing

719 perturbations may help efficiently capture the sensitivity field with a

720 smaller number of ensemble members. Its importance in practical

721 application prompted us to conduct another sensitivity analysis based on

722 100 ensemble members initialized from EnKF-based perturbations for

723 Shanshan. The settings are the same as in section 4 except that the

724 sensitivity field is derived from Eqs. (3) and (4) by using EnKF-based

725 perturbations instead of large-scale uncorrelated perturbations.

726 Figure 16 shows the TC forecast tracks initialized from EnKF-based

727 perturbations. Ensemble spreads of the forecasted TC central longitude and

728 latitude at 0000 UTC 17 September are 1.22 and 1.40 degree (a distance of

729 114 km and 156 km), respectively. They are larger than those in section 4

730 presumably due to the faster development of perturbations and the larger

40

731 displacement at the initial time. Figure 17 shows the sensitivity of TC

732 forecast latitude with respect to the vorticity field. The values at  =

733 0.8445 and = 0.5060 are larger than those at = 0.2879. Generally

734 speaking, the sensitivity is relatively strong north of 25N. In particular,

735 there is a pair pattern in the north of the initial TC position which is

736 inclined to the north with increasing altitude. These sensitivity signals are

737 quite different from those obtained in section 4, likely due to the property

738 of perturbations. Detailed discussion on this topic will be covered in future

739 works.

740 In order to verify that these sensitivity signals are relevant to the TC

741 motion, two ensemble experiments are performed as in sections 4c and 5c.

LAT 742 Initial vorticity perturbation with the magnitude of  is added

743 in the “POSITIVE” experiment, while the initial vorticity perturbation with

LAT 744 the magnitude of  is introduced to the other experiment

745 termed “NEGATIVE”. Here,   is the ensemble spread of the vorticity at

746 the initial time, and  is set to the constant value of 0.01 so that the

747 maximum change is about the size of the ensemble spread.

748 Fig. 18 shows that the changes of TC positions are consistent with

41

749 those obtained from the sensitivity field. The vortex displacement as a

750 response to the initial vorticity changes was located in the north-northeast

751 (south-southwest) direction as expected, while the distance of displacement

752 is slightly overestimated. It exhibits that 100 EnKF-based perturbations can

753 be utilized to detect the sensitivity of typhoon forecast positions through

754 TyPOS.

755

756 7. Concluding Remarks

757 Selection of the forecast metric is one of the critical issues in

758 sensitivity analyses associated with studies of TC motion. In this study, we

759 proposed a typhoon-position-oriented sensitivity analysis (TyPOS) for

760 ensemble forecast system. The sensitivity is interpreted as the slope of the

761 regression line (or its approximation) based on the ensemble simulation

762 between the initial perturbations and the resulted forecast scalar metric

763 representing the TC position.

764 As a first step, Typhoon Shanshan (2006) and Typhoon Dolphin

765 (2008) are used to demonstrate the applicability of TyPOS as a response to

766 large-scale initial perturbations. The sensitivity field of the TC central

42

767 latitude with respect to the vorticity field is maximized in the middle

768 troposphere. Interestingly, the sensitivity field is characterized by a dipole

769 pattern and the horizontally tilted patterns centered at the initial TC

770 position. The sensitivity signals related to the vorticity field in the middle

771 troposphere are stronger than those in other altitudes, while the sensitivity

772 related to the divergence field is less significant. This indicates that the

773 sensitivity signals shown by TyPOS are fairly consistent with the signals

774 obtained from existing methods, although there are still a few differences.

775 The sensitivity signals can be utilized to evaluate the changes in the

776 ensemble-mean TC position as a response to the changes in the initial

777 condition. They are also useful for determining which observations more

778 likely to objectively contribute to improving ensemble track forecasts. The

779 verification experiments confirm that sensitivity signals in TyPOS reflect

780 the displacement of the TC forecast position even at the verification time of

781 96 h.

782 As mentioned above, TyPOS is a useful method to resolve problems

783 in the choice of metrics. However, further improvements are needed

784 because TyPOS is still subject to a variety of uncertainties in specifying

43

785 initial perturbation magnitudes and patterns. In particular, limitations

786 resulting from the large-scale perturbations would make it difficult to

787 address the finer structure of the sensitivity field. Some noise arises as the

788 number of ensemble members decreases and also with a finer horizontal

789 grid spacing, although the t-test results serve as a tool to evaluate whether

790 the acquired signals are reliable or not.

791 The use of the EnKF-based perturbations may help efficiently

792 capture the sensitivity field with a smaller number of ensemble members

793 and reflect the sophisticated treatment of perturbation development through

794 a data assimilation system. Another sensitivity analysis based on 100

795 members initialized from EnKF-based perturbations is conducted for

796 Shanshan. It is encouraging to find that the ensemble-mean displacement of

797 TC forecast position is consistent with that expected by the sensitivity field.

798 However, the sensitivity for EnKF-based perturbations is clearly different

799 from that obtained for large-scale perturbations. Further study is required to

800 explain the factors leading to such a difference.

801 Other issues to be further investigated include the physical

802 interpretation of the signals, detailed comparison with other methodologies

44

803 and applications to real cases in which a bimodal distribution of TC

804 positions is obtained from ensemble runs. Despite these remaining

805 problems, the results shown here are meaningful because a distinctly

806 objective sensitivity analysis method for ensemble TC track forecasts is

807 proposed and verified. This method has potentials to contribute to further

808 reductions of ensemble forecast track errors in connection with targeted

809 observations, and to improved understanding of physical processes

810 associated with TC motion.

811

812 Acknowledgments.

813 Special thanks to J.-H. Chen, S.-G. Chen, G.-Y. Lien, M. Ueno, Y.-H.

814 Huang, N. Sugiura, and T. Unuma for their useful comments in developing

815 the current system. This work was funded by National Science Council

816 (NSC) grant 98-2111-M-002-008-MY3 and NSC100-2119-M-002-009-

817 MY3 in Taiwan. The figures were produced by the GFD-DENNOU Library.

45

818 REFERENCES

819 Aberson, S., 2003: Targeted observations to improve operational tropical cyclone track

820 forecast guidance. Mon. Wea. Rev., 131, 1613.

821 Aberson, S. and J. Franklin, 1999: Impact on hurricane track and intensity forecasts of

822 GPS dropwindsonde observations from the first-season flights of the NOAA

823 Gulfstream-IV jet aircraft. Bull. Amer. Meteor. Soc., 80 (3), 421-427.

824 Ancell, B. and G. Hakim, 2007: Comparing adjoint-and ensemble-sensitivity analysis

825 with applications to observation targeting. Mon. Wea. Rev., 135 (12), 4117-4134.

826 Aonashi, K. and H. Eito, 2011: Displaced ensemble variational assimilation method to

827 incorporate microwave imager brightness temperatures into a cloud-resolving

828 model. J. Meteor. Soc. Japan, 89 (3), 175-194.

829 Box, G. and M. Muller, 1958: A note on the generation of random normal deviates. The

830 Annals of Mathematical Statistics, 29 (2), 610-611.

831 Buizza, R., C. Cardinali, G. Kelly, and J. Th_epaut, 2007: The value of observations II:

832 The value of observations located in singular-vector-based target areas. Quart. J.

833 Roy. Meteor. Soc., 133, 1817-1832.

834 Chan, J., 2010: Movement of tropical cyclones. Global Perspectives on Tropical

835 Cyclones: From Science to Mitigation, World Scientific Publishing, 133-148.

836 Chen, J.-H., M. S. Peng, C. A. Reynolds, and C.-C. Wu, 2009: Interpretation of TC

837 forecast sensitivity from the SV perspective. J. Atmos. Sci. 66, 3383-3400.

838 Chen, S.-G., C.-C. Wu, J.-H. Chen, and K.-H. Chou, 2011: Validation and

839 interpretation of adjoint-derived sensitivity steering vector as targeted observation

840 guidance. Mon. Wea. Rev., 139, 1608-1625.

841 Chou, K. and C. Wu, 2008: Typhoon initialization in a mesoscale model-combination

46

842 of the bogused vortex and the dropwindsonde data in dotstar. Mon. Wea. Rev., 136

843 (3), 865-879.

844 Chou, K.-H., C.-C. Wu, P.-H. Lin, S. Aberson, M. Weissmann, F. Harnisch, and T.

845 Nakazawa, 2011: The impact of dropwindsonde observations on typhoon track

846 forecasts in DOTSTAR and T-PARC. Mon. Wea. Rev., 139 (6), 1728-1743.

847 Dudhia, J., 1989: Numerical study of convection observed during the winter monsoon

848 experiment using a mesoscale two-dimensional model. J. Atmos. Sci., 46 (20),

849 3077-3107.

850 Elsberry, R., 1995: Tropical cyclone motion. Global Perspectives on Tropical

851 Cyclones, World Meteorological Organization, 106-197.

852 Elsberry, R. and P. Harr, 2008: Tropical cyclone structure (TCS08) field experiment

853 science basis, observational platforms, and strategy. Asia-Pacific J. Atmos. Sci, 44

854 (3), 209-231.

855 Gombos, D. and J. Hansen, 2008: Potential vorticity regression and its relationship to

856 dynamical piecewise inversion. Mon. Wea. Rev., 136 (7), 2668-2682.

857 Gombos, D., R. Hoffman, and J. Hansen, 2012: Ensemble statistics for diagnosing

858 dynamics: Tropical cyclone track forecast sensitivities revealed by ensemble

859 regression. Mon. Wea. Rev., 140, 2647-2669.

860 Grell, G. and D. Devenyi, 2002: A generalized approach to parameterizing convection

861 combining ensemble and data assimilation techniques. Geophys. Res. Lett., 29,

862 doi:10.1029/2002GL015 311, 2002.

863 Hong, S.-Y., J. Dudhia, and S. Chen, 2004: A revised approach to ice microphysical

864 processes for the bulk parameterization of clouds and precipitation. Mon. Wea. Rev.,

865 132 (1), 103-120.

47

866 Hong, S.-Y. and J.-O. Lim, 2006: The WRF single-moment 6-class microphysics

867 scheme (WSM6). J. Korean Meteor. Soc, 42 (2), 129-151.

868 Hong, S.-Y., Y. Noh, and J. Dudhia, 2006: A new vertical diffusion package with an

869 explicit treatment of entrainment processes. Mon. Wea. Rev., 134 (9), 2318-2341.

870 Huang, Y.-H., M. Montgomery, and C.-C. Wu, 2012: Concentric eyewall formation in

871 typhoon sinlaku (2008). Part II: Axisymmetric dynamical processes. J. Atmos. Sci.,

872 69 (2), 662-674.

873 Ito, K., T. Kawabata, T. Kato, Y. Honda, Y. Ishikawa, and T. Awaji, 2013:

874 Simultaneous optimization of air-sea exchange coefficients and initial conditions

875 near a tropical cyclone with JNoVA. submitted to J. Meteorol. Soc. Jap., in minor

876 revision.

877 Japan Meteorological Agency, 2009: Annual Report on the Activities of the RSMC

878 Tokyo-Typhoon Center 2008. http://www.jma.go.jp/jma/jma-eng/jma-center/rsmc-

879 hp-pub-eg/AnnualReport/2008/Text/Text2008.pdf (accessed on June 20, 2012).

880 Japan Meteorological Agency, 2010: Annual Report on the Activities of the RSMC

881 Tokyo - Typhoon Center 2009. http://www.jma.go.jp/jma/jma-eng/jma-center/rsmc-

882 hp-pub-eg/AnnualReport/2009/Text/Text2009.pdf (accessed on December 4, 2011).

883 Kim, H., S. Kim, and B. Jung, 2011: Real-time adaptive observation guidance using

884 singular vectors for Typhoon Jangmi (200815) in T-PARC 2008. Wea. Fore., 26,

885 634-649.

886 Kim, H.-M. and B.-J. Jung, 2009: Singular vector structure and evolution of a

887 recurving tropical cyclone. Mon. Wea. Rev., 137 (2), 505-524.

888 Komaromi, W., S. Majumdar, and E. Rappin, 2011: Diagnosing initial condition

889 sensitivity of typhoon Sinlaku (2008) and hurricane Ike (2008). Mon. Wea. Rev.,

48

890 139 (10), 3224-3242.

891 Liu, C., Q. Xiao, and B. Wang, 2008: An ensemble-based four-dimensional variational

892 data assimilation scheme. Part I: Technical formulation and preliminary test. Mon.

893 Wea. Rev., 136 (9), 3363-3373.

894 Majumdar, S., S. Aberson, C. Bishop, R. Buizza, M. Peng, and C. Reynolds, 2006: A

895 comparison of adaptive observing guidance for atlantic tropical cyclones. Mon. Wea.

896 Rev., 134 (9), 2354-2372.

897 Majumdar, S., S. Chen, and C.Wu, 2011: Characteristics of ensemble transform

898 Kalman filter adaptive sampling guidance for tropical cyclones. Quart. J. Roy.

899 Meteor. Soc., 137, 503-520.

900 Martin, W. and M. Xue, 2006: Sensitivity analysis of convection of the 24 may 2002

901 IHOP case using very large ensembles. Mon. Wea. Rev., 134 (1), 192-207.

902 Matsumoto, M. and T. Nishimura, 1998: Mersenne twister: A 623-dimensionally

903 equidistributed uniform pseudo-random number generator. ACM Transactions on

904 Modeling and Computer Simulation, 8 (1), 3-30.

905 Meng, Z. and F. Zhang, 2007: Tests of an ensemble Kalman filter for mesoscale and

906 regional-scale data assimilation. Part II: Imperfect model experiments. Mon. Wea.

907 Rev., 135 (4), 1403-1423.

908 Meng, Z. and F. Zhang, 2008a: Tests of an ensemble Kalman filter for mesoscale and

909 regional-scale data assimilation. Part III: Comparison with 3DVAR in a real-data

910 case study. Mon. Wea. Rev., 136 (2), 522-540.

911 Meng, Z. and F. Zhang, 2008b: Tests of an ensemble Kalman filter for mesoscale and

912 regional-scale data assimilation. Part IV: Comparison with 3DVAR in a month-long

913 experiment. Mon. Wea. Rev., 136 (10), 3671-3682.

49

914 Menke, W., 1989: Geophysical data analysis: discrete inverse theory, 2nd edition, Vol.

915 45. Academic Press.

916 Mlawer, E., S. Taubman, P. Brown, M. Iacono, and S. Clough, 1997: Radiative transfer

917 for inhomogeneous atmospheres: RRTM, a validated correlated-k model for the

918 longwave. J. Geophys. Res., 102 (D14), 16 663-16 682, doi:10.1029/97JD00 237,

919 1997.

920 Mu, M., F. Zhou, and H.Wang, 2009: A method for identifying the sensitive areas in

921 targeted observations for tropical cyclone prediction: Conditional nonlinear optimal

922 perturbation. Mon. Wea. Rev., 137 (5), 1623-1639.

923 National Hurricane Center, 2012: National Hurricane Center Forecast Veri_cation.

924 Updated 1 March 2012, http://www.nhc.noaa.gov/veri_cation/ (accessed on 5 June

925 2012), updated 1 March 2012.

926 Peng, M., R. Maue, C. Reynolds, and R. Langland, 2007: Hurricanes Ivan, Jeanne,

927 Karl (2004) and mid-latitude trough interactions. Meteorol. Atmos. Phys., 97 (1),

928 221-237.

929 Peng, M. and C. Reynolds, 2006: Sensitivity of tropical cyclone forecasts as revealed

930 by singular vectors. J. Atmos. Sci., 63 (10), 2508-2528.

931 Reynolds, C., M. Peng, S. Majumdar, S. Aberson, C. Bishop, and R. Buizza, 2007:

932 Interpretation of adaptive observing guidance for atlantic tropical cyclones. Mon.

933 Wea. Rev., 135 (12), 4006-4029.

934 Reynolds, C., M.-S. Peng, and J.-H. Chen, 2009: Recurving tropical cyclones:

935 Singular vector sensitivity and downstream impacts. Mon. Wea. Rev., 137 (4), 1320-

936 1337.

937 Sugiura, N., 2010: A research on data assimilation methods towards the estimation and

50

938 the prediction of ocean variabilities on seasonal, interannual, and decadal

939 timescales. Ph.D. thesis, Kyoto University, 144 pp. (in Japanese).

940 Torn, R. and G. Hakim, 2008: Ensemble-based sensitivity analysis. Mon. Wea. Rev.,

941 136 (2), 663-677.

942 Torn, R. and G. Hakim, 2009: Initial condition sensitivity of western pacific

943 extratropical transitions determined using ensemble-based sensitivity analysis. Mon.

944 Wea. Rev., 137 (10), 3388-3406.

945 Weissmann, M., et al., 2011: The inuence of assimilating dropsonde data on typhoon

946 track and midlatitude forecasts. Mon. Wea. Rev., 139 (3), 908-920.

947 Wilks, D., 2005: Statistical methods in the atmospheric sciences, Second edition.

948 Academic Press, 648 pp.

949 Wu, C.-C. and K. Emanuel, 1993: Interaction of a baroclinic vortex with background

950 shear: Application to hurricane movement. J. Atmos. Sci., 50 (1), 62-76.

951 Wu, C.-C. and Y.-H. Kuo, 1999: Typhoons affecting Taiwan: Current understanding

952 and future challenges. Bull. Amer. Meteor. Soc., 80 (1), 67-80.

953 Wu, C.-C., and Coauthors, 2005: Dropwindsonde observations for typhoon

954 surveillance near the Taiwan Region (DOTSTAR). Bull. Amer. Meteor. Soc., 86 (6),

955 787–790.

956 Wu, C.-C., J.-H. Chen, P.-H. Lin, and K.-H. Chou, 2007: Targeted observations of

957 tropical cyclone movement based on the adjoint-derived sensitivity steering vector.

958 J. Atmos. Sci., 64 (7), 2611-2626.

959 Wu, C.-C., and Co-authors, 2009a: Intercomparison of targeted observation guidance

960 for tropical cyclones in the northwestern Pacific. Mon. Wea. Rev., 137 (8), 2471-

961 2492.

51

962 Wu, C.-C., S.-G. Chen, J.-H. Chen, K.-H. Chou, and P.-H. Lin, 2009b: Interaction of

963 Typhoon Shanshan (2006) with the midlatitude trough from both adjoint-derived

964 sensitivity steering vector and potential vorticity perspectives. Mon. Wea. Rev., 137

965 (3), 852-862.

966 Wu, C.-C., G.-Y. Lien, J.-H. Chen, and F. Zhang, 2010: Assimilation of tropical

967 cyclone track and structure based on the ensemble Kalman filter (EnKF). J. Atmos.

968 Sci., 67 (12),3806-3822.

969 Wu, C.-C., S.-G. Chen, C.-C. Yang, P.-H. Lin, and S. Aberson, 2012a: Potential

970 Vorticity Diagnosis of the Factors Affecting the Track of Typhoon Sinlaku (2008)

971 and the Impact from Dropwindsonde Data during T-PARC. Mon. Wea. Rev., 140,

972 2670-2688.

973 Wu, C.-C., Y.-H. Huang, and G.-Y. Lien, 2012b: Concentric Eyewall Formation in

974 Typhoon Sinlaku (2008). Part I: Assimilation of T-PARC Data Based on the

975 Ensemble Kalman Filter (EnKF). Mon. Wea. Rev., 140, 506-527.

976 Yamaguchi, M., T. Iriguchi, T. Nakazawa, and C.-C. Wu, 2009: An observing system

977 experiment for Typhoon Conson (2004) using a singular vector method and

978 DOTSTAR data. Mon. Wea. Rev., 137 (9), 2801-2816.

979 Yamaguchi, M., T. Nakazawa, and K. Aonashi (2012), Tropical cyclone track forecasts

980 using JMA model with ECMWF and JMA initial conditions, Geophys. Res. Lett.,

981 39, L09801, doi:10.1029/2012GL051473.

982 Yoden, S., 2007: Atmospheric predictability. J. Meteor. Soc. Japan, 85, 77-102.

983

52

984 Figure captions

985 1. Schematic illustration of the sensitivity with respect to the variable of single-

986 degree-of-freedom. Angle brackets denote the ensemble-mean values.

987 2. The domains used in this experiment. The dashed line in (a) indicates TC

988 position from JMA best track for typhoon Shanshan. Closed circles indicate TC

989 positions at time intervals of 24 hours from 0000Z 11 September. (b) Same as (a) but

990 for typhoon Dolphin. Closed circles indicate the TC positions every 24 hours from

991 0000 UTC 12 December 2008.

992 3. Time schedule of the sensitivity analysis for (a) typhoon Shanshan and (b)

993 typhoon Dolphin.

994 4. TC position from JMA best track for typhoon Shanshan plotted over simulated

995 tracks. Circles indicate the TC positions every 6 hours from 2100 UTC 10 September,

996 while solid ones indicate the TC positions every 24 hours from 0000 UTC 11

997 September. The TC symbols and thick solid lines indicate the ensemble-mean

998 simulated track from 0000 UTC 15 September to 0000 UTC 17 September. The gray

999 lines correspond to the tracks of each member. The cross marks (x) indicate the

1000 simulated positions of TC center at 0000 UTC 17 September.

1001 5. Ensemble-mean geopotential height (shadings) and horizontal wind (vectors) at

1002 (a)-(c) 300 hPa and (d)-(f) 500 hPa based on 800 member forecasts at (a)(d) 0000

1003 UTC 15 September, (b)(e) 0000 UTC 16 September and (c)(f) 0000 UTC 17

1004 September. Unit vectors for horizontal wind fields (m s-1) are shown on the rhs of each

1005 panel. The TC symbol indicates the ensemble-mean simulated position at each time.

1006 6. Ensemble-based sensitivity of typhoon Shanshan's central latitude with respect

1007 to the vorticity field (Unit in degree s). Circles in the panels indicate the results of

53

1008 significance t-test at each reduced grid point. The open TC symbols indicate the

1009 ensemble-mean TC positions at the initial time and closed TC symbols indicate those

1010 at the verification time. Left color bar indicates the values in the left column and right

1011 color bar indicates the values in the middle and right columns.

1012 7. Horizontal velocity perturbations relative to CTRL in (a) WTILT, (b) ETILT and

1013 (c) TROUGH. Unit vectors for horizontal wind fields (m s-1) are shown on the rhs of

1014 each panel. Vectors that are too small are not plotted. The open TC symbol indicates

1015 the position of typhoon Shanshan at the initial time, while the closed one indicates the

1016 ensemble-mean position at the verification time in the CTRL run.

1017 8. Time evolution of horizontal velocity perturbations on 500 hPa in ETILT_m001

1018 relative to CTRL_m001 at (a) 0, (b) 12, (c) 24 and (d) 48 h. Unit vectors for horizontal

1019 wind fields (m s-1) are shown on the rhs of each panel. Shadings indicate the

1020 geopotential height in CTRL_m001. The open TC symbol indicates the TC position in

1021 ETILT_m001, while the closed one indicates that of CTRL_m001.

1022 9. (a) Ensemble-mean TC track for CTRL (black), WTILT (red), ETILT (blue) and

1023 TROUGH (green). (b) Frequency distributions of the TC central latitude at the

1024 verification time of 48 h for the results of the CTRL, WTILT, ETILT and TROUGH

1025 runs. Vertical lines denote the ensemble-mean TC central latitude corresponding to the

1026 same line-color. (c) Ensemble-mean displacement of TC center position (relative to the

1027 CTRL run) at the verification time obtained from the nonlinear numerical model run.

1028 Horizontal and vertical axes indicate the displacement in longitude and latitude (Units

1029 in arc minute), respectively. (d) Same as (c), but for those evaluated by using TyPOS.

1030 10. TC position from JMA best track for Dolphin plotted over simulated tracks.

1031 Circles indicate the TC positions every 6 hours from 0600 UTC 11 December, while

54

1032 solid ones indicate the TC positions every 24 hours from 0000 UTC 12 December. The

1033 TC symbols and thick solid lines indicate the ensemble-mean simulated track from

1034 1200 UTC 11 December to 1200 UTC 15 December. The gray lines correspond to the

1035 tracks of each member. The cross marks (x) indicate the simulated positions of TC

1036 center at 1200 UTC 15 September.

1037 11. Same as Fig. 5 but for ensemble-mean geopotential height and horizontal wind

1038 around Dolphin on (a)(d) 1200 UTC 11 December, (b)(e) 1200 UTC 13 December and

1039 (c)(f) 1200 UTC 15 December.

1040 12. Same as Fig. 6 but for the sensitivity of Dolphin's longitude at the verification

1041 time of 24, 48 and 96 h.

1042 13. Same as Fig. 7 but for the verification experiment on the sensitivity for Dolphin:

1043 (a) P1, (b) P2, (c) N1, and (d) N2.

1044 14. (a) Ensemble-mean TC tracks of CTRL (black), P1 (red) and P2 (orange). (b)

1045 Same as (a) but for CTRL, N1 (cyan) and N2 (blue) experiments. (c) Frequency

1046 distributions of the TC central longitude at the verification time for CTRL, P1 and P2

1047 runs. (d) Same as (c) but for CTRL, N1 and N2 experiments. (e)(f) Same as Figs. 9c,d

1048 but for Dolphin at the verification time of 96 h.

1049 15. Sensitivity with respect to the vorticity field at  = 0.5060 using (a)-(c) 800

1050 (d)-(f) 400 (g)-(i) 200 (j)-(l) 100 ensemble members. The metric represents the TC

1051 central latitude at (a)(d)(g)(j) 12, (b)(e)(h)(k) 24 and (c)(f)(i)(l) 48 hours.

1052 16. Same as Fig. 6 but for the sensitivity analysis with a reduced-grid spacing of 270

1053 km.

1054 17. Same as Fig. 4 except that the ensemble simulations are initialized from EnKF-

1055 based perturbations.

55

1056 18. Same as Fig. 6 except that the ensemble simulations are initialized from EnKF-

1057 based perturbations.

1058 19. Same as Fig. 9 but for the POSITIVE and NEGATIVE experiments.

56

1059 1060 1061 Fig. 1. Schematic illustration of the sensitivity with respect to the variable 1062 of single-degree-of-freedom. Angle brackets denote the ensemble-mean 1063 values. 1064 1065

57

1066 1067

1068 1069 Fig. 2. The domains used in this experiment. The dashed line in (a) 1070 indicates TC position from JMA best track for typhoon Shanshan. Closed 1071 circles indicate TC positions at time intervals of 24 hours from 0000Z 11 1072 September. (b) Same as (a) but for typhoon Dolphin. Closed circles indicate 1073 the TC positions every 24 hours from 0000 UTC 12 December 2008. 1074

58

1075 1076 Fig. 3. Time schedule of the sensitivity analysis for (a) typhoon Shanshan 1077 and (b) typhoon Dolphin. 1078 1079 1080 1081

59

1082 1083 Fig. 4. TC position from JMA best track for typhoon Shanshan plotted over 1084 simulated tracks. Circles indicate the TC positions every 6 hours from 2100 1085 UTC 10 September, while solid ones indicate the TC positions every 24 1086 hours from 0000 UTC 11 September. The TC symbols and thick solid lines 1087 indicate the ensemble-mean simulated track from 0000 UTC 15 September 1088 to 0000 UTC 17 September. The gray lines correspond to the tracks of each 1089 member. The cross marks (x) indicate the simulated positions of TC center 1090 at 0000 UTC 17 September. 1091

60

1092 1093 1094 Fig. 5. Ensemble-mean geopotential height (shadings) and horizontal wind 1095 (vectors) at (a)-(c) 300 hPa and (d)-(f) 500 hPa based on 800 member 1096 forecasts at (a)(d) 0000 UTC 15 September, (b)(e) 0000 UTC 16 September 1097 and (c)(f) 0000 UTC 17 September. Unit vectors for horizontal wind fields 1098 (m s-1) are shown on the rhs of upper and lower panels. The TC symbol 1099 indicates the ensemble-mean simulated position at each time.

61

1100 1101 1102 Fig. 6. Ensemble-based sensitivity of typhoon Shanshan's central latitude 1103 with respect to the vorticity field (Unit in degree s). Circles in the panels 1104 indicate the results of significance t-test at each reduced grid point. The 1105 open TC symbols indicate the ensemble-mean TC positions at the initial 1106 time and closed TC symbols indicate those at the verification time. Left 1107 color bar indicates the values in the left column and right color bar 1108 indicates the values in the middle and right columns. 1109 1110

62

1111 1112 Fig. 7. Horizontal velocity perturbations relative to CTRL in (a) WTILT, 1113 (b) ETILT and (c) TROUGH. Unit vectors for horizontal wind fields (m s- 1114 1) are shown on the rhs of each panel. Vectors that are too small are not 1115 plotted. The open TC symbol indicates the position of typhoon Shanshan at 1116 the initial time, while the closed one indicates the ensemble-mean position 1117 at the verification time in the CTRL run. 1118

63

1119 1120 Fig. 8. Time evolution of horizontal velocity perturbations on 500 hPa in 1121 ETILT_m001 relative to CTRL_m001 at (a) 0, (b) 12, (c) 24 and (d) 48 h. 1122 Unit vectors for horizontal wind fields (m s-1) are shown on the rhs of each 1123 panel. Shadings indicate the geopotential height in CTRL_m001. The open 1124 TC symbol indicates the TC position in ETILT_m001, while the closed one 1125 indicates that of CTRL_m001.

64

1126 1127 Fig. 9. (a) Ensemble-mean TC track for CTRL (black), WTILT (red), 1128 ETILT (blue) and TROUGH (green). (b) Frequency distributions of the TC 1129 central latitude at the verification time of 48 h for the results of the CTRL, 1130 WTILT, ETILT and TROUGH runs. Vertical lines denote the ensemble- 1131 mean TC central latitude corresponding to the same line-color. (c) 1132 Ensemble-mean displacement of TC center position (relative to the CTRL 1133 run) at the verification time obtained from the nonlinear numerical model 1134 run. Horizontal and vertical axes indicate the displacement in longitude and 1135 latitude (Units in arc minute), respectively. (d) Same as (c), but for those 1136 evaluated by using TyPOS.

65

1137 1138 Fig. 10. TC position from JMA best track for Dolphin plotted over 1139 simulated tracks. Circles indicate the TC positions every 6 hours from 0600 1140 UTC 11 December, while solid ones indicate the TC positions every 24 1141 hours from 0000 UTC 12 December. The TC symbols and thick solid lines 1142 indicate the ensemble-mean simulated track from 1200 UTC 11 December 1143 to 1200 UTC 15 December. The gray lines correspond to the tracks of each 1144 member. The cross marks (x) indicate the simulated positions of TC center 1145 at 1200 UTC 15 September.

66

1146 1147 Fig. 11. Same as Fig. 5 but for ensemble-mean geopotential height and 1148 horizontal wind around Dolphin on (a)(d) 1200 UTC 11 December, (b)(e) 1149 1200 UTC 13 December and (c)(f) 1200 UTC 15 December.

67

1150 1151 Fig. 12. Same as Fig. 6 but for the sensitivity of Dolphin's longitude at the 1152 verification time of 24, 48 and 96 h. 1153 1154

68

1155 1156 Fig. 13. Same as Fig. 9 but for the verification experiment on the sensitivity 1157 for Dolphin: (a) P1, (b) P2, (c) N1, and (d) N2.

69

1158 1159 Fig. 14. (a) Ensemble-mean TC tracks of CTRL (black), P1 (red) and P2 1160 (orange). (b) Same as (a) but for CTRL, N1 (cyan) and N2 (blue) 1161 experiments. (c) Frequency distributions of the TC central longitude at the 1162 verification time for CTRL, P1 and P2 runs. (d) Same as (c) but for CTRL, 1163 N1 and N2 experiments. (e)(f) Same as Figs. 9c,d but for Dolphin at the 1164 verification time of 96 h.

70

1165 1166 Fig. 15. Sensitivity with respect to the vorticity field at  = 0.5060 using 1167 (a)-(c) 800 (d)-(f) 400 (g)-(i) 200 (j)-(l) 100 ensemble members. The metric 1168 represents the TC central latitude at (a)(d)(g)(j) 12, (b)(e)(h)(k) 24 and 1169 (c)(f)(i)(l) 48 hours.

71

1170 1171 Fig. 16. Same as Fig. 6 but for the sensitivity analysis with a reduced-grid 1172 spacing of 270 km. 1173 1174

72

1175 1176 1177 Fig. 17. Same as Fig. 4 except that the ensemble simulations are 1178 initialized from EnKF-based perturbations. 1179

73

1180 1181 Fig. 18. Same as Fig. 6 except that the ensemble simulations are 1182 initialized from EnKF-based perturbations. 1183 1184 1185 1186

74

1187 1188 1189 Fig. 19. Same as Fig. 9 but for the POSITIVE and NEGATIVE experiments. 1190

75