Advanced Search
Article Contents

A Timescale Decomposed Threshold Regression Downscaling Approach to Forecasting South China Early Summer Rainfall


doi: 10.1007/s00376-016-5251-7

  • A timescale decomposed threshold regression (TSDTR) downscaling approach to forecasting South China early summer rainfall (SCESR) is described by using long-term observed station rainfall data and NOAA ERSST data. It makes use of two distinct regression downscaling models corresponding to the interannual and interdecadal rainfall variability of SCESR. The two models are developed based on the partial least squares (PLS) regression technique, linking SCESR to SST modes in preceding months on both interannual and interdecadal timescales. Specifically, using the datasets in the calibration period 1915-84, the variability of SCESR and SST are decomposed into interannual and interdecadal components. On the interannual timescale, a threshold PLS regression model is fitted to interannual components of SCESR and March SST patterns by taking account of the modulation of negative and positive phases of the Pacific Decadal Oscillation (PDO). On the interdecadal timescale, a standard PLS regression model is fitted to the relationship between SCESR and preceding November SST patterns. The total rainfall prediction is obtained by the sum of the outputs from both the interannual and interdecadal models. Results show that the TSDTR downscaling approach achieves reasonable skill in predicting the observed rainfall in the validation period 1985-2006, compared to other simpler approaches. This study suggests that the TSDTR approach, considering different interannual SCESR-SST relationships under the modulation of PDO phases, as well as the interdecadal variability of SCESR associated with SST patterns, may provide a new perspective to improve climate predictions.
  • 加载中
  • Aligo E. A., W. A. Gallus Jr., and M. Segal, 2009: On the impact of WRF model vertical grid resolution on Midwest summer rainfall forecasts. Wea. Forecasting, 24, 575- 594.10.1175/2008WAF2007101.1f4ca9824-eb91-4f75-b465-c9176184dbd0b7f8bed1d1974b9fe11868b43ff7e71ehttp%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F2009WtFor..24..575Arefpaperuri:(f20e2db75277de4beae88fb0c0a52bae)http://adsabs.harvard.edu/abs/2009WtFor..24..575ANot Available
    Butler N. A., M. C. Denham, 2000: The peculiar shrinkage properties of partial least squares regression. Journal of the Royal Statistical Society:Series B (Statistical Methodology), 62, 585- 593.10.1111/1467-9868.00252682cef5330f28fc9d5488afd7194f93ehttp%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1111%2F1467-9868.00252%2Fcitedbyhttp://onlinelibrary.wiley.com/doi/10.1111/1467-9868.00252/citedbyPartial least squares regression has been widely adopted within some areas as a useful alternative to ordinary least squares regression in the manner of other shrinkage methods such as principal components regression and ridge regression. In this paper we examine the nature of this shrinkage and demonstrate that partial least squares regression exhibits some undesirable properties.
    Cao X., S. F. Chen, G. H. Chen, W. Chen, and R. G. Wu, 2015: On the weakened relationship between spring Arctic Oscillation and following summer tropical cyclone frequency over the western north Pacific: A comparison between 1968-1986 and 1989-2007. Adv. Atmos. Sci.,32, 1319-1328, doi: 10.1007/ s00376-015-4256-y.10.1007/s00376-015-4256-y297f2111-d849-44bc-a77f-58cb426b214d056ad5c2dd49c962e47973e1dfb48aa6http%3A%2F%2Flink.springer.com%2F10.1007%2Fs00376-015-4256-yrefpaperuri:(8d747c268a2aaa412a81659f68c3efdc)http://d.wanfangdata.com.cn/Periodical/dqkxjz-e201510001This study documents a weakening of the relationship between the spring Arctic Oscillation(AO) and the following summer tropical cyclone(TC) formation frequency over the eastern part(150?–180?E) of the western North Pacific(WNP). The relationship is strong and statistically significant during 1968–1986,but becomes weak during 1989–2007. The spring AOrelated SST,atmospheric dynamic,and thermodynamic conditions are compared between the two epochs to understand the possible reasons for the change in the relationship. Results indicate that the spring AO leads to an El Ni ?no-like SST anomaly,lower-level anomalous cyclonic circulation,upper-level anomalous anticyclonic circulation,enhanced ascending motion,and a positive midlevel relative humidity anomaly in the tropical western–central Pacific during 1968–1986,whereas the AOrelated anomalies in the above quantities are weak during 1989–2007. Hence,the large-scale dynamic and thermodynamic anomalies are more favorable for TC formation over the eastern WNP during 1968–1986 than during 1989–2007.
    Cao X., S. F. Chen, G. H. Chen, and R. G. Wu, 2016: Intensified impact of northern tropical Atlantic SST on tropical cyclogenesis frequency over the western North Pacific after the late 1980s. Adv. Atmos. Sci., doi: 10.1007/s00376-016-5206-z.10.1007/s00376-016-5206-z7040ef1381e9ca882072db3932015a29http%3A%2F%2Fwww.cnki.com.cn%2FArticle%2FCJFDTotal-DQJZ201608002.htmhttp://www.cnki.com.cn/Article/CJFDTotal-DQJZ201608002.htmPrevious studies suggest that spring SST anomalies over the northern tropical Atlantic(NTA) affect the tropical cyclone(TC) activity over the western North Pacific(WNP) in the following summer and fall. The present study reveals that the connection between spring NTA SST and following summer–fall WNP TC genesis frequency is not stationary. The influence of spring NTA SST on following summer–fall WNP TC genesis frequency is weak and insignificant before, but strong and significant after, the late 1980 s. Before the late 1980 s, the NTA SST anomaly-induced SST anomalies in the tropical central Pacific are weak, and the response of atmospheric circulation over the WNP is not strong. As a result, the connection between spring NTA SST and following summer–fall WNP TC genesis frequency is insignificant in the former period. In contrast,after the late 1980 s, NTA SST anomalies induce pronounced tropical central Pacific SST anomalies through an Atlantic–Pacific teleconnection. Tropical central Pacific SST anomalies further induce favorable conditions for WNP TC genesis,including vertical motion, mid-level relative humidity, and vertical zonal wind shear. Hence, the connection between NTA SST and WNP TC genesis frequency is significant in the recent period. Further analysis shows that the interdecadal change in the connection between spring NTA SST and following summer–fall WNP TC genesis frequency may be related to the climatological SST change over the NTA region.
    Chan J. C. L., W. Zhou, 2005: PDO, ENSO and the early summer monsoon rainfall over south China. Geophys. Res. Lett., 32, L08810.
    Chen S.-F., W. Chen, B. Yu, and H.-F. Graf, 2013a: Modulation of the seasonal footprinting mechanism by the boreal spring Arctic Oscillation. Geophys. Res. Lett., 40, 6384-6389, doi: 10.1002/2013GL058628.10.1002/2013GL058628943bdb818c1c623cc03afdd69984a8b0http%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1002%2F2013GL058628%2Fabstracthttp://onlinelibrary.wiley.com/doi/10.1002/2013GL058628/abstractstudies suggest that wintertime North Pacific Oscillation (NPO) is able to force an El Ni09o event during the following winter via the seasonal footprinting mechanism (SFM). In this study, we present evidence that springtime Arctic Oscillation (AO) has a significant modulation effect on the connection between the NPO and El Ni09o. When the spring AO is positive, a positive wintertime NPO can result in significant El Ni09o-like warming anomalies via the SFM. However, when the spring AO is negative, the connection of the NPO and the El Ni09o is not robust at all. Thus, the phase of spring AO should be taken into account when using NPO as a predictor for El Ni09o. Further analysis reveals that the mechanism of the AO modulation may be through changing the underlying SST footprinting over the subtropical northeastern Pacific.
    Chen S. F., B. Yu, and W. Chen, 2014: An analysis on the physical process of the influence of AO on ENSO. Climate Dyn., 42, 973- 989.10.1007/s00382-012-1654-zf9267c6aed2e083ad58a362574e6e2cbhttp%3A%2F%2Fcpfd.cnki.com.cn%2FArticle%2FCPFDTOTAL-ZGQX201307001007.htmhttp://cpfd.cnki.com.cn/Article/CPFDTOTAL-ZGQX201307001007.htm正The influence of the spring AO on ENSO has been demonstrated in several recent studies.This analysis further explores the physical process of the influence of AO on ENSO using the NCEP/NCAR reanalysis data over the period 1958-2010.We focus on the formation of the westerly wind burst in the tropical western Pacific,and examine the evolution and formation of the atmospheric circulation,atmospheric heating,and SST
    Chen S. F., B. Yu, and W. Chen, 2015a: An interdecadal change in the influence of the spring Arctic Oscillation on the subsequent ENSO around the early 1970s. Climate Dyn., 44, 1109- 1126.10.1007/s00382-014-2152-236b577af-ea32-4d06-b1bb-3da50f5eaf32c2f19db9632fe7a1539af37cd1387ba6http%3A%2F%2Flink.springer.com%2F10.1007%2Fs00382-014-2152-2refpaperuri:(79aea9aad505b85ecce6ee63b7671cd6)http://link.springer.com/10.1007/s00382-014-2152-2Previous studies suggested that the springtime Arctic Oscillation (AO) influences the El Niño-Southern Oscillation (ENSO) outbreak in the following winter. Using the HadISST, HadSLP2r, ERSSTv3b and NCEP-NCAR reanalysis data for the period 1948-2012, this analysis further reveals that the AO-ENSO relationship experienced a pronounced interdecadal shift. The spring AO influence on the subsequent ENSO is weak before 1970; while the influence becomes strong and statistically significant in the 1970s and 1980s. We then compare the spring AO associated circulation, SST and precipitation anomalies between the PRE (1949-1968) and POST (1970-1989) epochs to explore this interdecadal change of the AO-ENSO relationship. The spring AO-related anomalies of atmospheric circulation over the North Pacific mid-latitudes, cyclonic circulation over the subtropical western-central Pacific, and westerly winds in the tropical western-central Pacific are found to be stronger in the POST epoch than in the PRE epoch. The intensity of spring Pacific synoptic-scale eddy activity is seen to experience a significant interdecadal change around the early-1970s from a weak regime to a strong regime. Thus the strength of synoptic-scale eddy feedback to the low frequency flow becomes stronger after 1970. In the POST epoch, the strong synoptic-scale eddy feedback provides a favorable condition for the formulation of the spring AO-related cyclonic circulation and westerly wind anomalies over the western North Pacific. The tropical SST, precipitation and atmospheric circulation anomalies sustain and develop from spring to winter through the positive Bjerknes feedback, leading to an El Niño-like warming in the tropical central-eastern Pacific in the following winter.
    Chen S. F., W. Chen, and R. G. Wu, 2015b: An interdecadal change in the relationship between boreal spring Arctic Oscillation and the East Asian Summer Monsoon around the early 1970s. J.Climate, 28, 1527- 1542.10.1175/JCLI-D-14-00409.1822d14f9-9c6e-4b40-9bc1-7e7cd068f6eb4e77bf7338125bdbf8d0de436b235320http%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F2015JCli...28.1527Crefpaperuri:(f21d5ab300fdd7b9f1cc662aaa384809)http://adsabs.harvard.edu/abs/2015JCli...28.1527CNot Available
    Chen W., J. Feng, and R. G. Wu, 2013b: Roles of ENSO and PDO in the link of the East Asian Winter Monsoon to the following Summer Monsoon. J.Climate, 26, 622- 635.10.1175/JCLI-D-12-00021.1c3184660ce1313978d3d2a3777696772http%3A%2F%2Fonlinelibrary.wiley.com%2Fresolve%2Freference%2FXREF%3Fid%3D10.1175%2FJCLI-D-12-00021.1http://onlinelibrary.wiley.com/resolve/reference/XREF?id=10.1175/JCLI-D-12-00021.1Abstract The present study investigates the roles of El Ni09o–Southern Oscillation (ENSO) and the Pacific decadal oscillation (PDO) in the relationship between the East Asian winter monsoon (EAWM) and the following East Asian summer monsoon (EASM). The variability of the EAWM is divided into an ENSO-related part named EAWM EN and an ENSO-unrelated part named EAWM res . Corresponding to a weak EAWM EN , an anomalous low-level anticyclone forms over the western North Pacific (WNP) and persists from winter to the following summer. This anticyclone enhances southerlies over the coast of East Asia in summer. Hence, a weak EAWM EN tends to be followed by a strong EASM and vice versa. As such, a link is established between the EAWM EN and the EASM. The persistence of this WNP anticyclone may be mainly attributed to the sea surface temperature anomalies associated with the ENSO-related EAWM part in the tropical Indian Ocean and the extratropical North Pacific. In contrast, corresponding to a weak EAWM res , the anomalous WNP anticyclone is only seen in winter, and there is no obvious relationship between the EAWM res and the following EASM. Therefore, the observed EAWM–EASM relationship is dominated by the winter monsoon variability associated with ENSO. It is found that the EAWM EN –EASM relationship is modulated by the PDO. There tends to be a much stronger EASM after a weak EAWM EN during the positive PDO phases than during the negative PDO phases.
    Collischonn W., R. Haas, I. Andreolli, and C. E. M. Tucci, 2005: Forecasting River Uruguay flow using rainfall forecasts from a regional weather-prediction model. J. Hydrol., 305, 87- 98.10.1016/j.jhydrol.2004.08.0280ef2e009c06fc54755ace86f08573272http%3A%2F%2Fwww.sciencedirect.com%2Fscience%2Farticle%2Fpii%2FS0022169404004093http://www.sciencedirect.com/science/article/pii/S0022169404004093The use of quantitative rainfall forecasts as input to a rainfall-runoff model, thereby extending the lead-time of flow forecasts, is relatively new. This paper presents results from a study in which real-time river flow forecasts were calculated for the River Uruguay basin lying within southern Brazil, using a method based on observed rainfall, quantitative forecasts of rainfall given by a regional numerical weather-prediction model, and rainfall-runoff simulation by a distributed hydrological model. The performance of discharge forecasts was evaluated over a continuous 167-day period and from one selected flood event, using rainfall forecasts at three spatial resolutions. The performance of these forecasts was also compared with that of forecasts obtained (a) by assuming that no further rain would fall, and (b) by assuming that rainfall forecasts were equal to the rainfall actually recorded, this representing a surrogate for 'perfect' rainfall forecasts. The results show that for the basin considered, there is plenty of scope for improving usefulness of rainfall forecasts.
    Ding Y. H., Z. Y. Wang, and Y. Sun, 2008: Inter-decadal variation of the summer precipitation in East China and its association with decreasing Asian summer monsoon. Part I: Observed evidences. Int. J. Climatol., 28, 1139- 1161.10.1002/joc.1615f6a9cdd8-fc22-4e0e-b778-b8a996cd66589cdf407839f313cb5748e3fed4f537e1http%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1002%2Fjoc.1615%2Ffullrefpaperuri:(f3f65327f855f4e6f32a79087d79643f)http://onlinelibrary.wiley.com/doi/10.1002/joc.1615/fullThe bonding in the molecule ion VO(H2O)(5)(2+) is described in terms of molecular orbitals. In particular, the most significant feature of the electronic structure of VO2+ seems to be the existence of considerable oxygen to vanadium pi-bonding. A molecular orbital energy level scheme is estimated which is able to account for both the "crystal field" and the "charge transfer" spectra of VO(H2O)(5)(2+) and related vanadyl complexes. The paramagnetic resonance g factors and the magnetic susceptibilities of vanadyl complexes are discussed.
    Duan W. S., L. Y. Song, Y. Li, and J. Y. Mao, 2013: Modulation of PDO on the predictability of the interannual variability of early summer rainfall over south China. J. Geophys. Res., 118, 13 008- 13 021.10.1002/2013JD0198627b930d35de0fe59ac68e763497f4ed00http%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1002%2F2013JD019862%2Ffullhttp://onlinelibrary.wiley.com/doi/10.1002/2013JD019862/fullstudy investigates the modulation of the Pacific Decadal Oscillation (PDO) on the predictability of interannual early summer south China rainfall (SCR) using high-quality station rainfall data. Of particular interest is the difference in impact between negative and positive phases of the PDO on the predictability of interannual early summer SCR. A clear difference in the correlation between the interannual early summer SCR and the preceding sea surface temperature (SST) over the Pacific Ocean appears in negative and positive phases of the PDO. In the negative PDO phase, the correlation between interannual early summer SCR and SST is dominated by a pattern with significant negative correlations in the subtropical western North Pacific and southeast Pacific and significant positive correlations in the tropical central Pacific. However, in the positive PDO phase, significant positive correlations are observed in the tropical eastern Pacific. It is found that, for each PDO phase, the preceding SST anomalies in some regions in the Pacific may act as predictors of the interannual early summer SCR. As such, a two-regime regression model for the relationship between interannual early summer SCR and preceding SST anomalies is established based on the negative and positive PDO phases using respective multiple linear regression models. Results suggest that the interannual early SCR is more predictable in PDO positive phase than in negative phase. It offers a support for the argument that a segmented statistical forecasting approach associated with the decadal modulation effect of the coupled ocean atmospheric mode should be adopted to forecast the early summer SCR.
    Eitzen Z. A., D. A. Randall, 1999: Sensitivity of the simulated Asian summer monsoon to parameterized physical processes. J. Geophys. Res., 104, 12 177- 12 191.10.1029/1999JD90016860195a75363df201137a774d1548680ahttp%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1029%2F1999JD900168%2Ffullhttp://onlinelibrary.wiley.com/doi/10.1029/1999JD900168/fullA study of the sensitivity of the simulated Asian summer monsoon to changes in general circulation model formulation is reported. The baseline version of the model fails to realistically simulate the precipitation, wind, and temperature fields. In one experiment the stratiform cloud parameterization was changed from a simple large-scale saturation scheme to a scheme that prognostically determines cloud water, cloud ice, and rain. In a second experiment a parameter that relates the cumulus mass flux to the cumulus kinetic energy was altered so as to increase the convective adjustment time. These changes in the stratiform and cumuliform cloud parameterizations significantly improve the simulations of the precipitation and upper level wind fields, respectively.
    Gao H., Y. G. Wang, and J. H. He, 2006: Weakening significance of ENSO as a predictor of summer precipitation in China. Geophys. Res. Lett., 33(9),L09807, doi: 10.1029/2005 GL025511.10.1029/2005GL025511c9813bf6f06a6ccd8e7d91e1ea9d15aehttp%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1029%2F2005GL025511%2Fabstracthttp://onlinelibrary.wiley.com/doi/10.1029/2005GL025511/abstract[1] The interdecadal variation of the relationship between ENSO and summer precipitation in China has been examined based on observed monthly rainfall data and NOAA ERSST data from 1951 to 2003. Results show that the relation has weakened during the past two decades, and the significance of ENSO as a predictor has also decreased. An evident example is that before the late 1970s, when above-normal (below-normal) SST appears over the Niño-3 or Niño-4 regions in previous winters, more (less) summer rainfall will often be found in North China and south of Yangtze River valley, less (more) rainfall appears along the Huaihe River valley, and the Chinese Meiyu will be later (earlier). However, all of these conclusions should be adopted carefully after the 1980s because of the feeble relation between ENSO and summer precipitation in China. This weakening relationship has increased the difficulty of summer rainfall prediction in China.
    Gershunov A., T. P. Barnett, 1998: Interdecadal modulation of ENSO teleconnections. Bull. Amer. Meteor. Soc., 79, 2715- 2726.e8627613-1da6-4c27-bad7-44b28d3a9800fdadbf2409f92113315adbf05b428686http%3A%2F%2Fwww.bioone.org%2Fservlet%2Flinkout%3Fsuffix%3Di1533-6085-38-1-45-Gershunov1%26dbid%3D16%26doi%3D10.2181%252F1533-6085%282005%29038%5B0045%253AHWDSHI%5D2.0.CO%253B2%26key%3D10.1175%252F1520-0477%281998%290792.0.CO%253B2refpaperuri:(c6d9dd5f86db24ab8d9b239e673d1506)http://xueshu.baidu.com/s?wd=paperuri%3A%28c6d9dd5f86db24ab8d9b239e673d1506%29&filter=sc_long_sign&tn=SE_xueshusource_2kduw22v&sc_vurl=http%3A%2F%2Fwww.bioone.org%2Fservlet%2Flinkout%3Fsuffix%3Di1533-6085-38-1-45-Gershunov1%26dbid%3D16%26doi%3D10.2181%252F1533-6085%282005%29038%5B0045%253AHWDSHI%5D2.0.CO%253B2%26key%3D10.1175%252F1520-0477%281998%290792.0.CO%253B2&ie=utf-8&sc_us=10019848668784770153
    Gong D.-Y., C.-H. Ho, 2002: Shift in the summer rainfall over the Yangtze River valley in the late 1970s. Geophys. Res. Lett.,29, 78-1-78-4.10.1029/2001GL014523bcfe793dac639818d834184f6971eb7chttp%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1029%2F2001GL014523%2Fcitedbyhttp://onlinelibrary.wiley.com/doi/10.1029/2001GL014523/citedbyThe summer rainfall over the middle‐lower valley of the Yangtze River and over the whole eastern China experienced a notable regime shift in about 1979. This change is consistent with a simultaneous jump‐like change in the 500 hPa geopotential height (Φ500) over the northern Pacific. The rainfall over the Yangtze River valley is closely related to the Φ500 averaged over the area 20°–25°N, 125°–140°E, with a correlation coefficient of 0.66 for the period 1958–1999. Since 1980, the subtropical northwestern Pacific high (SNPH) has enlarged, intensified, and extended southwestward. The changes in the SNPH are strongly associated with the variations of the sea surface temperatures (SSTs) of the eastern tropical Pacific and tropical Indian Ocean. The anomalies of these SSTs, responsible primarily for the shift of the summer rainfall over the Yangtze River through the changes in SNPH, precede the Φ500 signals with different leading times.
    Grotch S. L., M. C. MacCracken, 1991: The use of general circulation models to predict regional climatic change. J.Climate, 4, 286- 303.10.1175/1520-0442(1991)004<0286:TUOGCM>2.0.CO;2058152d99cb35f6093af62701bcb1f0ahttp%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F1991JCli....4..286Ghttp://adsabs.harvard.edu/abs/1991JCli....4..286GEquilibrium simulations using the best-available general circulation models to estimate the sensitivity of the climate to a doubling of the atmospheric carbon dioxide concentration are in broad general agreement that the global annual average surface air temperature would increase 2.5 to 4.5 K. However, at finer spatial scales, the range of changes in temperature and precipitation predicted by different computer models is much broader. Many shortcomings are also apparent in the model simulations of the present climate, indicating that further model improvements are needed to achieve reliable regional and seasonal projections of the future climatic conditions.
    Guo Y., J. P. Li, and Y. Li, 2012: A time-scale decomposition approach to statistically downscale summer rainfall over North China. J.Climate, 25, 572- 591.10.1175/JCLI-D-11-00014.12584b5136156b95b4938fd31cbb4d376http%3A%2F%2Fonlinelibrary.wiley.com%2Fresolve%2Freference%2FADS%3Fid%3D2012JCli...25..572Ghttp://onlinelibrary.wiley.com/resolve/reference/ADS?id=2012JCli...25..572GNot Available
    Haenlein M., A. M. Kaplan, 2004: A beginner's guide to partial least squares analysis. Understanding Statistics, 3, 283- 297.10.1207/s15328031us0304_465cd3f1ba1db4b13b6e8e6c684755e1ehttp%3A%2F%2Fwww.tandfonline.com%2Fdoi%2Fabs%2F10.1207%2Fs15328031us0304_4http://www.tandfonline.com/doi/abs/10.1207/s15328031us0304_4Since the introduction of covariance-based structural equation modeling (SEM) by Joreskog in 1973, this technique has been received with considerable interest among empirical researchers. However, the predominance of LISREL, certainly the most well-known tool to perform this kind of analysis, has led to the fact that not all researchers are aware of alternative techniques for SEM, such as partial least squares (PLS) analysis. Therefore, the objective of this article is to provide an easily comprehensible introduction to this technique, which is particularly suited to situations in which constructs are measured by a very large number of indicators and where maximum likelihood covariance-based SEM tools reach their limit. Because this article is intended as a general introduction, it avoids mathematical details as far as possible and instead focuses on a presentation of PLS, which can be understood without an in-depth knowledge of SEM.
    Helland, I. S., 1988: On the structure of partial least squares regression. Communications in Statistics-Simulation and Computation, 17, 581- 607.10.1080/036109188088126819669d59127c8c40d9ad5063808ef4609http%3A%2F%2Fwww.ams.org%2Fmathscinet-getitem%3Fmr%3D955342http://www.ams.org/mathscinet-getitem?mr=955342We prove that the two algorithms given in the literature for partial least squares regression are equivalent, and use this equivalence to give an explicit formula for the resulting prediction equation. This in turn is used to investigate the regression method from several points of view. Its relation to principal component regression is clearified, and some heuristic arguments are given to explain why partial least squares regression often needs fewer factors to give its optimal prediction.
    Huang R. H., Y. F. Wu, 1989: The influence of ENSO on the summer climate change in China and its mechanism. Adv. Atmos. Sci.,6, 21-32, doi: 10.1007/BF02656915.10.1007/BF0265691569409ca5d9a91750ecdc40755db2445fhttp%3A%2F%2Fonlinelibrary.wiley.com%2Fresolve%2Freference%2FXREF%3Fid%3D10.1007%2FBF02656915http://www.cnki.com.cn/Article/CJFDTotal-DQJZ198901001.htmThe influence of ENSO on the summer climate change in China and its mechanism from the observed data is discussed. It is discovered that in the developing stage of ENSO, the SST in the western tropical Pacific is colder in summer, the convective activities may be weak around the South China Sea and the Philippines. As a consequence, the subtropical high shifted southward. Therefore, a drought may be caused in the Indo-China peninsula and in the South China. Moreover, in midsummer the subtropical high is weak over the Yangtze River valley and Huaihe River valley, and the flood may be caused in the area from the Yangtze River valley to Huaihe River valley. On the contrary, in the decaying stage of ENSO. the convective activities may be strong around the Philippines, and the subtropical high shifted northward, a drought may be caused in the Yangtze River valley and Huaihe River valley.
    Huang R. H., J. L. Chen, G. Huang, and Q. L. Zhang, 2006: The quasi-biennial oscillation of summer monsoon rainfall in China and its cause. Chinese Journal of Atmospheric Sciences, 30, 545- 560. (in Chinese)10.1063/1.3187794e6bacce6b991f7ab182204cdf6916915http%3A%2F%2Fen.cnki.com.cn%2FArticle_en%2FCJFDTOTAL-DQXK200604000.htmhttp://en.cnki.com.cn/Article_en/CJFDTOTAL-DQXK200604000.htmThe observed data of precipitation at 160 observational stations of China,the ERA-40 reanalysis data and the Empirical Orthogonal Function(EOF) and the entropy spectral analysis methods are applied to analyze the interannual variations of summer(JuneAugust) rainfall in China and water vapor transport fluxes over East Asia.The results show that there is an obvious oscillation with a period of two or three years,i.e.,the quasi-biennial oscillation,in the interannual variations of summer monsoon rainfall in China,especially in the eastern and southern parts of China including South China,the Yangtze River valley and the Huaihe River valley and North China.And it is also shown that this oscillation is closely associated with the quasi-biennial oscillation in the interannual variations of the water vapor transport fluxes by summer monsoon flow over East Asia.Furthermore,the interannual variations of sea temperature in the surface and subsurface of the tropical western Pacific are analyzed by using the sea surface temperature(SST) data from the NCEP/NCAR reanalysis dataset and the sea temperature data in the subsurface of the western Pacific along 137° E from Japan Meteorological Agency,respectively.And it is revealed that there is also a significant quasi-biennial oscillation in the interannual variations of thermal state of the tropical western Pacific.In this paper,the correlative and composite analysis methods are applied to discuss the influence of the quasi-biennial oscillation of thermal state of the tropical western Pacific on summer rainfall in China and water vapor transport over East Asia,and it is shown that the quasi-biennial oscillation in the interannual variations of thermal state of the tropical western Pacific has a great impact on the East Asian summer monsoon and the water vapor transport driven by the monsoon flow.Besides,the influence of the quasi-biennial oscillation in the interannual variations of thermal state of the tropical western Pacific on the quasi-biennual oscillation in the interannual variations of the summer monsoon rainfall in China is simply discussed by using the teleconnection theory of the East Asia/Pacific(EAP) pattern. From the above-mentioned analyses,the physical mechanism of the quasi-biennial oscillation of summer rainfall in China may be summarized as follows: If the thermal state of the tropical western Pacific is in a warming state during a winter,then the convective activities will be intensified around the Philippines in the following spring and summer,which can cause weak summer monsoon rainfall in the Yangtze River and the Huaihe River valleys through the EAP pattern teleconnection.And due to the intensification of the convective activities around the Philippines,a strong convergence of atmospheric circulation will appear over the tropical western Pacific.This will trigger a strong upwelling in the tropical western Pacific.As a consequence,the thermal state of this region will turn into a cooling one in the following winter.On the other hand,since the tropical western Pacific will in a cooling state during the following winter,the convective activities will weaken around the Philippines in the spring and summer of the third year,which can cause strong summer monsoon rainfall in the Yangtze River and the Huaihe River valleys through the EAP pattern teleconnection.And due to the weakening of the convective activities around the Philippines,a divergence of atmospheric circulation will appear over the tropical western Pacific in the spring and summer of the third year.As a consequence,the thermal state of these ocean regions will again turn into a warming one in the winter of the third year.
    Kalnay, E., Coauthors, 1996: The NCEP/NCAR 40-year reanalysis project. Bull. Amer. Meteor. Soc.,77, 437-472, doi: 10.1175/1520-0477(1996)077<0437:TNYRP>2.0.CO;2.f539a4fb-a013-4942-ac7e-7f15017eedac23d674534321ec5c56bf181fd85f5561http%3A%2F%2Fwww.bioone.org%2Fservlet%2Flinkout%3Fsuffix%3Di1536-1098-69-2-93-Kalnay1%26dbid%3D16%26doi%3D10.3959%252F1536-1098-69.2.93%26key%3D10.1175%252F1520-0477%281996%29077%3C0437%253ATNYRP%3E2.0.CO%253B2refpaperuri:(fe1c070047a030c900beb40441caee5a)http://xueshu.baidu.com/s?wd=paperuri%3A%28fe1c070047a030c900beb40441caee5a%29&filter=sc_long_sign&tn=SE_xueshusource_2kduw22v&sc_vurl=http%3A%2F%2Fwww.bioone.org%2Fservlet%2Flinkout%3Fsuffix%3Di1536-1098-69-2-93-Kalnay1%26dbid%3D16%26doi%3D10.3959%252F1536-1098-69.2.93%26key%3D10.1175%252F1520-0477%281996%29077%253C0437%253ATNYRP%253E2.0.CO%253B2&ie=utf-8&sc_us=5666421642967837950
    Lau K. M., H. Y. Weng, 2001: Coherent modes of global SST and summer rainfall over China: An assessment of the regional impacts of the 1997-98 El Niño. J.Climate, 14, 1294- 1308.
    Li Y., I. Smith, 2009: A statistical downscaling model for southern Australia winter rainfall. J.Climate, 22, 1142- 1158.10.1175/2008JCLI2160.149dac67b26047bbadbda431935934f49http%3A%2F%2Fwww.cabdirect.org%2Fabstracts%2F20093130914.htmlhttp://www.cabdirect.org/abstracts/20093130914.htmlAbstract A technique for obtaining downscaled rainfall projections from climate model simulations is described. This technique makes use of the close association between mean sea level pressure (MSLP) patterns and rainfall over southern Australia during winter. Principal components of seasonal mean MSLP anomalies are linked to observed rainfall anomalies at regional, gridpoint, and point scales. A maximum of four components is sufficient to capture a relatively large fraction of the observed variance in rainfall at most locations. These are used to interpret the MSLP patterns from a single climate model, which has been used to simulate both present-day and future climate. The resulting downscaled values provide 1) a closer representation of the observed present-day rainfall than the raw climate model values and 2) alternative estimates of future changes to rainfall that arise owing to changes in mean MSLP. While decreases are simulated for later this century (under a single emissions scenario), the downscaled values, in percentage terms, tend to be less.
    Liu Y., K. Fan, 2012a: Improve the prediction of summer precipitation in the Southeastern China by a hybrid statistical downscaling model. Meteor. Atmos. Phys., 117, 121- 134.10.1007/s00703-012-0201-03b10a69e5d538270c2e5b63745a05d3bhttp%3A%2F%2Fwww.springerlink.com%2Fcontent%2Fj01j040h343755l4%2Fhttp://www.springerlink.com/content/j01j040h343755l4/We attempt to apply year-to-year increment prediction to develop an effective statistical downscaling scheme for summer (JJA, June–July–August) rainfall prediction at the station-to-station scale in Southeastern China (SEC). The year-to-year increment in a variable was defined as the difference between the current year and the previous year. This difference is related to the quasi-biennial oscillation in interannual variations in precipitation. Three predictors from observations and six from three general circulation models (GCMs) outputs of the development of a European multi-model ensemble system for seasonal to interannual prediction (DEMETER) project were used to establish this downscaling model. The independent sample test and the cross-validation test show that the downscaling scheme yields better predicted skill for summer precipitation at most stations over SEC than the original DEMETER GCM outputs, with greater temporal correlation coefficients and spatial anomaly correlation coefficients, as well as lower root-mean-square errors.
    Liu Y., K. Fan, 2012b: Prediction of spring precipitation in China using a downscaling approach. Meteor. Atmos. Phys., 118: 79- 93.10.1007/s00703-012-0202-z23f15e41dbc18be5f31e74c2f848dccdhttp%3A%2F%2Fwww.springerlink.com%2Fcontent%2F65n6v84880u43274%2Fhttp://www.springerlink.com/content/65n6v84880u43274/The aim of this paper is to use a statistical downscaling model to predict spring precipitation over China based on a large-scale circulation simulation using Development of a European Multi-model Ensemble System for Seasonal to Interannual Prediction (DEMETER) General Circulation Models (GCMs) from 1960 to 2001. A singular value decomposition regression analysis was performed to establish the link between the spring precipitation and the large-scale variables, particularly for the geopotential height at 500 hPa and the sea-level pressure. The DEMETER GCM predictors were determined on the basis of their agreement with the reanalysis data for specific domains. This downscaling scheme significantly improved the predictability compared with the raw DEMETER GCM output for both the independent hindcast test and the cross-validation test. For the independent hindcast test, multi-year average spatial correlation coefficients (CCs) increased by at least ~30 % compared with the DEMETER GCMs' precipitation output. In addition, the root mean-square errors (RMSEs) decreased more than 35 % compared with the raw DEMETER GCM output. For the cross-validation test, the spatial CCs increased to greater than 0.9 for most of the individual years, and the temporal CCs increased to greater than 0.3 (95 % confidence level) for most regions in China from 1960 to 2001. The RMSEs decreased significantly compared with the raw output. Furthermore, the preceding predictor, the Arctic Oscillation, increased the predicted skill of the downscaling scheme during the spring of 1963.
    Liu Y., K. Fan.2014: An application of hybrid downscaling model to forecast summer precipitation at stations in China. Atmospheric Research, 143: 17- 30.10.1016/j.atmosres.2014.01.024ad98db1666e420c17cdb9ef765daa9bdhttp%3A%2F%2Fwww.sciencedirect.com%2Fscience%2Farticle%2Fpii%2FS0169809514000581http://www.sciencedirect.com/science/article/pii/S0169809514000581A pattern prediction hybrid downscaling method was applied to predict summer (June ulyugust) precipitation at China 160 stations. The predicted precipitation from the downscaling scheme is available one month before. Four predictors were chosen to establish the hybrid downscaling scheme. The 500-hPa geopotential height (GH5) and 850-hPa specific humidity (q85) were from the skillful predicted output of three DEMETER (Development of a European Multi-model Ensemble System for Seasonal to Interannual Prediction) general circulation models (GCMs). The 700-hPa geopotential height (GH7) and sea level pressure (SLP) were from reanalysis datasets. The hybrid downscaling scheme (HD-4P) has better prediction skill than a conventional statistical downscaling model (SD-2P) which contains two predictors derived from the output of GCMs, although two downscaling schemes were performed to improve the seasonal prediction of summer rainfall in comparison with the original output of the DEMETER GCMs. In particular, HD-4P downscaling predictions showed lower root mean square errors than those based on the SD-2P model. Furthermore, the HD-4P downscaling model reproduced the China summer precipitation anomaly centers more accurately than the scenario of the SD-2P model in 1998. A hybrid downscaling prediction should be effective to improve the prediction skill of summer rainfall at stations in China.
    Mantua N. J., S. R. Hare, 2002: The Pacific decadal oscillation. Journal of Oceanography, 58, 35- 44.10.1023/A:1015820616384d98cc047-1c9c-40e2-ba2f-315e88bc521aslarticleid_14125678fa094039baed87220ef6b563ee94fe4http%3A%2F%2Flink.springer.com%2F10.1023%2FA%3A1015820616384refpaperuri:(7ee73282fc24a7b1d3afc7232a8a436b)http://link.springer.com/10.1023/A:1015820616384<a name="Abs1"></a>The <i>Pacific Decadal Oscillation</i> (PDO) has been described by some as a long-lived El Ni?o-like pattern of Pacific climate variability, and by others as a blend of two sometimes independent modes having distinct spatial and temporal characteristics of North Pacific sea surface temperature (SST) variability. A growing body of evidence highlights a strong tendency for PDO impacts in the Southern Hemisphere, with important surface climate anomalies over the mid-latitude South Pacific Ocean, Australia and South America. Several independent studies find evidence for just two full PDO cycles in the past century: &#8220;cool&#8221; PDO regimes prevailed from 1890&#8211;1924 and again from 1947&#8211;1976, while &#8220;warm&#8221; PDO regimes dominated from 1925&#8211;1946 and from 1977 through (at least) the mid-1990's. Interdecadal changes in Pacific climate have widespread impacts on natural systems, including water resources in the Americas and many marine fisheries in the North Pacific. Tree-ring and Pacific coral based climate reconstructions suggest that PDO variations&#8212;at a range of varying time scales&#8212;can be traced back to at least 1600, although there are important differences between different proxy reconstructions. While 20th Century PDO fluctuations were most energetic in two general periodicities&#8212;one from 15-to-25 years, and the other from 50-to-70 years&#8212;the mechanisms causing PDO variability remain unclear. To date, there is little in the way of observational evidence to support a mid-latitude coupled air-sea interaction for PDO, though there are several well-understood mechanisms that promote multi-year persistence in North Pacific upper ocean temperature anomalies.
    Mantua N. J., S. R. Hare, Y. Zhang, J. M. Wallace, and R. C. Francis, 1997: A Pacific interdecadal climate oscillation with impacts on salmon production. Bull. Amer. Meteor. Soc., 78, 1069- 1079.9a9ae4973967cb1c5f5681c069ca5d08http%3A%2F%2Ficesjms.oxfordjournals.org%2Fexternal-ref%3Faccess_num%3D10.1175%2F1520-0477%281997%290782.0.CO%3B2%26link_type%3DDOIhttp://icesjms.oxfordjournals.org/external-ref?access_num=10.1175/1520-0477(1997)0782.0.CO;2&amp;link_type=DOI
    Mao J. Y., J. C. L. Chan, and G. X. Wu, 2011: Interannual variations of early summer monsoon rainfall over South China under different PDO backgrounds. Int. J. Climatol., 31, 847- 862.10.1002/joc.2129a5d18d77cad2b0b1b796eafd55db0e47http%3A%2F%2Fso.med.wanfangdata.com.cn%2FViewHTML%2FPeriodicalPaper_JJ026388074.aspxhttp://so.med.wanfangdata.com.cn/ViewHTML/PeriodicalPaper_JJ026388074.aspxNot Available
    Martin G. M., 1999: The simulation of the Asian summer monsoon, and its sensitivity to horizontal resolution, in the UK meteorological office unified model. Quart. J. Roy. Meteor. Soc., 125, 1499- 1525.10.1002/qj.4971255570350e3aab2-8305-4b35-9b51-f1c2b1cbe93bcbb5f867cc1bd56af81a6065c136ba1bhttp%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1002%2Fqj.49712555703%2Fcitedbyrefpaperuri:(e7f9270219bf0356a2b8e9b6e4a8116f)http://onlinelibrary.wiley.com/doi/10.1002/qj.49712555703/citedbyNot Available
    Nitta T., Z.-Z. Hu, 1996: Summer climate variability in China and its association with 500 hPa height and tropical convection. J. Meteor. Soc. Japan Ser.II, 74, 425- 445.10.1175/1520-0469(1996)053<2283:OTFOLI>2.0.CO;289b69c2c64c12a09c2d165ea0548b543http%3A%2F%2Fci.nii.ac.jp%2Fnaid%2F10013127230%2Fhttp://ci.nii.ac.jp/naid/10013127230/This paper is concernd with interannual and interdecadal variabilities of summer rainfall and temperature patterns in China and their association with 500 hPa height in the Northern Hemisphere (NH), tropical convective activities and global sea surface temperature anomaly (SSTA). The temporal evolutions and spatial structures of interannual variation of summer (JJA) rainfall and temperature from 1951 to 1994 over China are revealed through EOF analysis. The spatial pattern of EOF1 for rainfall (EOF1.R) is dominated by a maximum over the middle-lower reaches of the Yangtze River, and a large negative value region in the middle reach of the Yellow River is also obvious. The spatial pattern of EOF1 for temperature (EOF1.T) reflects coherent variations over most regions of China, and it is dominated by a maximum over the middle-lower reaches of the Yangtze River. Linear increase and decrease trends are found in the time coefficients of EOF1.R and EOF1.T, respectively. The quasi-biennial oscillation (QBO) signal is also strong after the middle of the 1970's in repect of their time coefficients. The coupled patterns of rainfall and temperature are picked up through the singular value decomposition (SVD) analysis. The spatial patterns and their temporal evolutions of SVD1 for rainfall (SVD1.R) and SVD1 for temperature (SVD1.T) are quite similar to those of EOF1.R and EOF1.T. There is an abrupt change in the middle 1970's in the time coefficients of SVD2.R and SVD2.T. The variations of summer rainfall and temperature coupled patterns in China are closely connected with the 500 hPa height anomaly over the Northern Hemisphere (NH). The Pacific-Japan (PJ) and Eurasia (EU) teleconnection patterns play a very important role in the spatial patterns of SVD1.R and SVD1.T, especially in the East Asia monsoon region along the middle-lower reaches of the Yangtze River. The abrupt change of China summer climate in the middle 1970's is related with the intensification and southerly location of the western Pacific subtropical high and also the geopotential height changes over Eurasia and in the regions to the north of the Japan Sea in 1977 or 1978. Correlations between the summer rainfall and temperature coupled patterns and monthly-averaged out-going longwave radiation (OLR) and high-cloud amount (HCA) data are significant with the PJ teleconnection pattern. There exist positive correlations between the coupled patterns and sea surface temperature anomaly (SSTA) in the North Pacific and the tropical western Paciftc. A comparison study shows that there are coherent variations between summer rainfall in the middle-lower reaches of the Yangtze River and in the western part of Japan. It is also demonstrated that there are close correlations between the summer temperature variations in China and in Japan.
    Power S., T. Casey, C. Folland , A. Colman, and V. Mehta, 1999: Inter-decadal modulation of the impact of ENSO on Australia. Climate Dyn., 15, 319- 324.10.1007/s003820050284e3eeaf4aa6c1967d9320d0c332cf45e3http%3A%2F%2Fonlinelibrary.wiley.com%2Fresolve%2Freference%2FXREF%3Fid%3D10.1007%2Fs003820050284http://onlinelibrary.wiley.com/resolve/reference/XREF?id=10.1007/s003820050284The success of an ENSO-based statistical rainfall prediction scheme and the influence of ENSO on Australia are shown to vary in association with a coherent, inter-decadal oscillation in surface temperature over the Pacific Ocean. When this Inter-decadal Pacific Oscillation ( IPO ) raises temperatures in the tropical Pacific Ocean, there is no robust relationship between year-to-year Australian climate variations and ENSO. When the IPO lowers temperature in the same region, on the other hand, year-to-year ENSO variability is closely associated with year-to-year variability in rainfall, surface temperature, river flow and the domestic wheat crop yield. The contrast in ENSO influence between the two phases of the IPO is quite remarkable. This highlights exciting new avenues for obtaining improved climate predictions.
    Sahai A. K., A. M. Grimm, V. Satyan, and G. B. Pant, 2003: Long-lead prediction of Indian summer monsoon rainfall from global SST evolution. Climate Dyn., 20, 855- 863.10.1007/s00382-003-0306-8617c33bdc4174df96007d1d932d5ddaahttp%3A%2F%2Flink.springer.com%2Farticle%2F10.1007%2Fs00382-003-0306-8http://link.springer.com/article/10.1007/s00382-003-0306-8Not Available
    Smith T. M., R. W. Reynolds, 2004: Improved extended reconstruction of SST (1854-1997). J.Climate, 17, 2466- 2477.88ba6adeb2694186f76e7004c999bad4http%3A%2F%2Ficesjms.oxfordjournals.org%2Fexternal-ref%3Faccess_num%3D10.1175%2F1520-0442%282004%290172.0.CO%3B2%26link_type%3DDOIhttp://icesjms.oxfordjournals.org/external-ref?access_num=10.1175/1520-0442(2004)0172.0.CO;2&amp;link_type=DOI
    Smith T. M., R. W. Reynolds, T. C. Peterson, and J. Lawrimore, 2008: Improvements to NOAA's historical merged land-ocean surface temperature analysis (1880-2006). J.Climate, 21, 2283- 2296.
    Smoliak B. V., J. M. Wallace, M. T. Stoelinga, and T. P. Mitchell, 2010: Application of partial least squares regression to the diagnosis of year-to-year variations in Pacific Northwest snowpack and Atlantic hurricanes. Geophys. Res. Lett., 37,L03801, doi: 10.1029/2009GL041478.10.1029/2009GL0414785a4d06f0-846a-4a7f-a69f-d025c4878a4d1623ab6a2c82fa51c4390cc572f806bdhttp%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1029%2F2009GL041478%2Fabstractrefpaperuri:(b776174fab01b5cb805acba6de5d94ee)http://onlinelibrary.wiley.com/doi/10.1029/2009GL041478/abstractApplication of the method of partial least squares (PLS) regression to geophysical data is illustrated with two cases: (1) finding sea level pressure patterns over the North Pacific associated with dynamically-induced winter-to-winter variations in snowpack in the Cascade mountains of western Washington state and (2) finding patterns of sea surface temperature over the tropical oceans that modulate Atlantic hurricane activity on a year-to-year basis. In both examples two robust patterns in the “predictor field” are identified that, in combination, account for over half the variance in the target time series.
    Sun J. Q., H. J. Wang, 2012: Changes of the connection between the summer North Atlantic Oscillation and the East Asian summer rainfall. J. Geophys. Res., 117,D08110, doi: 10.1029/2012JD017482.10.1029/2012JD017482cdc87c4bd8d7211e711abb6f80c190c0http%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1029%2F2012JD017482%2Fabstracthttp://onlinelibrary.wiley.com/doi/10.1029/2012JD017482/abstract[1] In this study, the relationship between the summer North Atlantic Oscillation (SNAO) and the East Asian summer rainfall was statistically diagnosed based on the European Centre for Medium-Range Weather Forecasts (ECMWF) 40-year and interim reanalysis data (ERA-40 and ERA-Interim) as well as precipitation data from the Global Precipitation Climatology Centre (GPCC). The results show that the decadal change of the SNAO pattern around the late 1970s significantly enhanced its connection with summer rainfall over central and northern East Asia. Over the period before the late 1970s, the SNAO-related circulations were dominant over the North Atlantic region. Consequently, there was a weak connection between the SNAO and the East Asian summer rainfall. However, over the period after the late 1970s, the SNAO pattern experienced a decadal change, with the southern center shifting eastward. Such changes in the SNAO pattern can alter the stationary wave activity over the Eurasian Continent, producing an anomalous meridional dipole pattern over East Asia. This dipole pattern can then change the divergence circulation, vertical motion, water vapor, and total cloud cover, which would consequently provide beneficial conditions for more (less) summer rainfall over central (northern) East Asia in a positive (negative)-phase SNAO year.
    Sun J. Q., H. P. Chen, 2012: A statistical downscaling scheme to improve global precipitation forecasting. Meteor. Atmos. Phys., 117, 87- 102.10.1007/s00703-012-0195-73101cf5175529a48404fd5354da1f3f1http%3A%2F%2Flink.springer.com%2Farticle%2F10.1007%2Fs00703-012-0195-7http://link.springer.com/article/10.1007/s00703-012-0195-7Based on hindcasts obtained from the “Development of a European Multimodel Ensemble system for seasonal to inTERannual prediction” (DEMETER) project, this study proposes a statistical downscaling (SD) scheme suitable for global precipitation forecasting. The key idea of this SD scheme is to select the optimal predictors that are best forecast by coupled general circulation models (CGCMs) and that have the most stable relationships with observed precipitation. Developing the prediction model and further making predictions using these predictors can extract useful information from the CGCMs. Cross-validation and independent sample tests indicate that this SD scheme can significantly improve the prediction capability of CGCMs during the boreal summer (June–August), even over polar regions. The predicted and observed precipitations are significantly correlated, and the root-mean-square-error of the SD scheme-predicted precipitation is largely decreased compared with the raw CGCM predictions. An inter-model comparison shows that the multi-model ensemble provides the best prediction performance. This study suggests that combining a multi-model ensemble with the SD scheme can improve the prediction skill for precipitation globally, which is valuable for current operational precipitation prediction.
    Tao S. Y., 1987: A review of recent research on the East Asian summer monsoon in China. J. Meteor. Soc.Japan, 70, 373- 396.7a7cf2cfdb1d11184ad32b44ecf07d62http%3A%2F%2Fci.nii.ac.jp%2Fnaid%2F10012388648http://ci.nii.ac.jp/naid/10012388648A review of recent research on the East Asian summer monsoon in China TAO S. Y. Monsoon Meteorology, 1987
    Thompson D. W. J., J. M. Wallace, 1998: The Arctic oscillation signature in the wintertime geopotential height and temperature fields. Geophys. Res. Lett., 25, 1297- 1300.10.1029/98GL00950adf244c1165dc0c5b3e8ecc1d4c5e7fehttp%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1029%2F98GL00950%2Fpdfhttp://onlinelibrary.wiley.com/doi/10.1029/98GL00950/pdfThe leading empirical orthogonal function of the wintertime sea-level pressure field is more strongly coupled to surface air temperature fluctuations over the Eurasian continent than the North Atlantic Oscillation (NAO). It resembles the NAO in many respects; but its primary center of action covers more of the Arctic, giving it a more zonally symmetric appearance. Coupled to strong fluctuations at the 50-hPa level on the intraseasonal, interannual, and interdecadal time scales, this rctic Oscillation (AO) can be interpreted as the surface signature of modulations in the strength of the polar vortex aloft. It is proposed that the zonally asymmetric surface air temperature and mid-tropospheric circulation anomalies observed in association with the AO may be secondary baroclinic features induced by the land-sea contrasts. The same modal structure is mirrored in the pronounced trends in winter and springtime surface air temperature, sea-level pressure, and 50-hPa height over the past 30 years: parts of Eurasia have warmed by as much as several K, sea-level pressure over parts of the Arctic has fallen by 4 hPa, and the core of the lower stratospheric polar vortex has cooled by several K. These trends can be interpreted as the development of a systematic bias in one of the atmosphere's dominant, naturally occurring modes of variability.
    Torrence C., P. J. Webster, 1999: Interdecadal changes in the ENSO-monsoon system. J.Climate, 12, 2679- 2690.0dae6c1bb1cfb867465b338a7f18621dhttp%3A%2F%2Fwww.bioone.org%2Fservlet%2Flinkout%3Fsuffix%3Dbibr23%26dbid%3D16%26doi%3D10.3969%252Fj.issn.1674-764x.2010.04.009%26key%3D10.1175%252F1520-0442%281999%290122.0.CO%253B2http://xueshu.baidu.com/s?wd=paperuri%3A%283fc276c036c796d554505961ca20e343%29&filter=sc_long_sign&tn=SE_xueshusource_2kduw22v&sc_vurl=http%3A%2F%2Fwww.bioone.org%2Fservlet%2Flinkout%3Fsuffix%3Dbibr23%26dbid%3D16%26doi%3D10.3969%252Fj.issn.1674-764x.2010.04.009%26key%3D10.1175%252F1520-0442%281999%290122.0.CO%253B2&ie=utf-8&sc_us=13828971752612169262
    Wang, B., LinHo, Y. S. Zhang, M. M. Lu, 2004: Definition of South China Sea monsoon onset and commencement of the East Asia summer monsoon. J.Climate, 17, 699- 710.10.1175/2932.176cdca49cd0f55dcf3fe9e51d2dfc7a6http%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F2004jcli...17..699whttp://adsabs.harvard.edu/abs/2004jcli...17..699wABSTRACT The climatological mean summer monsoon onset in the South China Sea (SCS) is a remarkably abrupt event. However, defining onset dates for individual years is noticeably controversial. The controversies and complications arise primarily from a number of factors: the intermittent southward intrusion of cold fronts into the northern SCS, the bogus onset in late April before the establishment of tropical monsoons over Indochina, and active intraseasonal oscillation. In this paper, a simple yet effective index, USCS, the 850-hPa zonal winds averaged over the central SCS (5° 15°N and 110° 120°E), is proposed for objectively defining the SCS monsoon onset. This onset index depicts not only the sudden establishment of the tropical southwesterly monsoon over the SCS but also the outbreak of the rainy season in the central-northern SCS.In this paper the East Asian summer monsoon (EASM) is defined as the broadscale summer monsoon over East Asia and the western North Pacific region (0° 40°N, 100° 140°E). It is shown that the seasonal transition of EASM can be objectively determined by the principal component of the dominant empirical orthogonal mode of the 850-hPa zonal winds, UEOF1. It is found that the local index USCS represents UEOF1 extremely well; thus, it can be used to determine both the SCS onset and the commencement of the broadscale EASM. The result suggests that the SCS summer monsoon onset indeed signifies the beginning of the summer monsoon over East Asia and the adjacent western Pacific Ocean. Evidence is also provided to show the linkage between the two salient phases of EASM: the local onset of the SCS monsoon and the local onset of the mei-yu (the rainy season in the Yangtze River and Huai River basin and southern Japan).
    Wang B., R. G. Wu, and X. Fu, 2000: Pacific-east Asian teleconnection: How does ENSO affect east Asian climate? J.Climate, 13, 1517- 1536.f39da437-71e7-4c4d-aadf-fae7bdf913a15dc62d69fc115b6acbc741e3911fc4f7http%3A%2F%2Fwww.bioone.org%2Fservlet%2Flinkout%3Fsuffix%3Di1551-5036-22-3-625-Wang1%26dbid%3D16%26doi%3D10.2112%252F04-0156.1%26key%3D10.1175%252F1520-0442%282000%290132.0.CO%253B2refpaperuri:(c25afe041658a4f704de554223c4d38e)http://xueshu.baidu.com/s?wd=paperuri%3A%28c25afe041658a4f704de554223c4d38e%29&filter=sc_long_sign&tn=SE_xueshusource_2kduw22v&sc_vurl=http%3A%2F%2Fwww.bioone.org%2Fservlet%2Flinkout%3Fsuffix%3Di1551-5036-22-3-625-Wang1%26dbid%3D16%26doi%3D10.2112%252F04-0156.1%26key%3D10.1175%252F1520-0442%282000%290132.0.CO%253B2&ie=utf-8&sc_us=573464951031400502
    Wang H. J., 2001: The weakening of the Asian monsoon circulation after the end of 1970's. Adv. Atmos. Sci.,18(3), 376-386, doi: 10.1007/BF02919316.10.1007/BF0291931655186f36fb4d3628dc9d5fb9107f180bhttp%3A%2F%2Fonlinelibrary.wiley.com%2Fresolve%2Freference%2FXREF%3Fid%3D10.1007%2FBF02919316http://en.cnki.com.cn/Article_en/CJFDTotal-DQJZ200103004.htmThe transition of the global atmospheric circulation in the end of 1970's can clearly be detected in the atmospheric temperature, wind velocity, and so on. Wavelet analysis reveals that the temporal scale of this change is larger than 20 years. Studies in this work indicate that the trend of the transition over the mid-latitude Asia is opposite to that of global average for some variables at the middle troposphere. Another finding of this research is that the African-Asian monsoon circulation is weaker and the trade wind over the tropical eastern Pacific is weaker as well after this transition. Such a signal may be found in the summer precipitation over China as well.
    Wang H. J., 2002: The instability of the East Asian summer monsoon-ENSO relations. Adv. Atmos. Sci.,19(1), 1-11, doi: 10.1007/s00376-002-0029-5.10.1007/s00376-002-0029-583afa2a8-1414-49a8-ab33-2a0523507e600b0b8196ff36c1ce99353d997ae7053bhttp%3A%2F%2Flink.springer.com%2F10.1007%2Fs00376-002-0029-5refpaperuri:(cd03980142f2285999c71ee67cf31b4e)http://en.cnki.com.cn/Article_en/CJFDTOTAL-DQJZ200201000.htmThe instability in the relation between the East Asian summer monsoon and the ENSO cycle in the long-term variation is found through this research. By instability, we mean that high inter-relation exists in some periods but low inter-relation may appear in some other periods. It is reveals that the interannual variation of the summer atmospheric circulation during the ' high correlation' periods (HCP) is significantly different from that during the ' low correlation' periods (LCP). Larger interannual variability is found during HCP for trade wind over the tropical eastern Pacific of the Southern Hemisphere, the low-level air temperature over the tropical Pacific, the subtropical high pressure systems in the two hemispheres, and so on. The correlation between summer rainfall over China and ENSO is as well different between HCP and LCP.
    Wang L., W. Chen, R. H. Huang, 2008: Interdecadal modulation of PDO on the impact of ENSO on the east Asian winter monsoon. Geophys. Res. Lett., 35,L20702, doi: 10.1029/2008 GL035287.10.1029/2008GL0352872a10502ba5c1942b1d4ce693b9fbb478http%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1029%2F2008GL035287%2Fpdfhttp://onlinelibrary.wiley.com/doi/10.1029/2008GL035287/pdfThe interdecadal modulation of the Pacific Decadal Oscillation (PDO) upon the impact of the El Niño-Southern Oscillation (ENSO) on the east Asian winter monsoon (EAWM) is investigated. When the PDO is in its high phase, there is no robust relationship between ENSO and EAWM on the interannual timescale. When the PDO is in its low phase, ENSO exerts strong impact on the EAWM, with robust and significant low-level temperature changes occurring over east Asia. The contrast in ENSO's influence between the two phases of the PDO is quite remarkable, which urges that the phase of the PDO should be taken into account in the ENSO-based prediction of wintertime climate over east Asia. This modulation may be accounted for by the change in the low-latitude Pacific-east Asian teleconnection and in the response of midlatitude geopotential height to ENSO over Northwest Pacific.
    Wilby R. L., 1998: Statistical downscaling of daily precipitation using daily airflow and seasonal teleconnection indices. Climate Research, 10, 163- 178.10.3354/cr010163b7013121b44b6c91c88a1317d994d18dhttp%3A%2F%2Fwww.researchgate.net%2Fpublication%2F250221625_Statistical_downscaling_of_daily_precipitation_using_daily_airflow_and_seasonal_teleconnection_indiceshttp://www.researchgate.net/publication/250221625_Statistical_downscaling_of_daily_precipitation_using_daily_airflow_and_seasonal_teleconnection_indicesCiteSeerX - Scientific documents that cite the following paper: Statistical downscaling of daily precipitation using daily airflow and seasonal teleconnection indices
    Wu R. G., B. Wang, 2002: A contrast of the East Asian summer monsoon-ENSO relationship between 1962-77 and 1978-93. J.Climate, 15, 3266- 3279.10.1175/1520-0442(2002)0152.0.CO;2669caff818ce3d677d91fc1a568dafcdhttp%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F2002JCli...15.3266Whttp://adsabs.harvard.edu/abs/2002JCli...15.3266WUsing station rainfall data and the NCEP-NCAR reanalysis, the authors investigate changes in the interannual relationship between the east Asian summer monsoon (EASM) and El Ni01±o-Southern Oscillation (ENSO) in the late 1970s, concurrent with the Pacific climate shift. The present study focuses on decaying phases of ENSO because changes in developing phases of ENSO are less significant. Remarkable changes are found in the summer rainfall anomaly in northern China and Japan. From pre- to postshift period, the summer rainfall anomaly in eastern north China during decaying phases of El Ni01±o changed from above to below normal, whereas that in central Japan changed from negative to normal. Consistent with this, the barotropic anticyclonic anomaly over the Japan Sea changed to cyclonic; the associated anomalous winds changed from southerly to northerly over the Yellow Sea-northeastern China and from northeasterly to northwesterly over central Japan. The change in the ENSO-related east Asian summer circulation anomaly is attributed to changes in the location and intensity of anomalous convection over the western North Pacific (WNP) and India. After the late 1970s, the WNP convection anomaly is enhanced and shifted to higher latitudes due to increased summer mean SST in the Philippine Sea. This induces an eastward shift of an anomalous low pressure from east Asia to the North Pacific along 3000°-4500°N during decaying phases of El Ni01±o. Thus, anomalous winds over northeastern China and Korea switch from southeasterly to northeasterly. Before the late 1970s, an anomalous barotropic anticyclone develops over east Asia and anomalous southerties prevail over northeastern China during decaying phases of El Ni01±o. This may relate to anomalous Indian convection through a zonal wave pattern along 3000°-5000°N. After the late 1970s, anomalous Indian convection weakens, which reduces the impact of the Indian convection on the EASM.
    Wu R.G., S. Yang, Z. P. Wen, G. Huang, and K. M. Hu, 2012: Interdecadal change in the relationship of southern China summer rainfall with tropical Indo-Pacific SST. Theor. Appl. Climatol., 108, 119- 133.10.1007/s00704-011-0519-4b9cd6d864ba793790d7b1fee224f3742http%3A%2F%2Fonlinelibrary.wiley.com%2Fresolve%2Freference%2FADS%3Fid%3D2012ThApC.108..119Whttp://onlinelibrary.wiley.com/resolve/reference/ADS?id=2012ThApC.108..119WThe present study investigates the interdecadal change in the relationship between southern China (SC) summer rainfall and tropical Indo-Pacific sea surface temperature (SST). It is found that the pattern of tropical Indo-Pacific SST anomalies associated with SC summer rainfall variability tends to be opposite between the 1950-1960s and the 1980-1990s. Above-normal SC rainfall corresponds to warmer SST in the tropical southeastern Indian Ocean (SEIO) and cooler SST in the equatorial central Pacific (ECP) during the 1950-1960s but opposite SST anomalies in these regions during the 1980-1990s. A pronounced difference is also found in anomalous atmospheric circulation linking SEIO SST and SC rainfall between the two periods. In the 1950-1960s, two anomalous vertical circulations are present between ascent over SEIO and ascent over SC, with a common branch of descent over the South China Sea that is accompanied by an anomalous low-level anticyclone. In the 1980-1990s, however, a single anomalous vertical circulation directly connects ascent over SC to descent over SEIO. The change in the rainfall-SST relationship is likely related to a change in the magnitude of SEIO SST forcing and a change in the atmospheric response to the SST forcing due to different mean states. A larger SEIO SST forcing coupled with a stronger and more extensive western North Pacific subtropical high in recent decades induce circulation anomalies reaching higher latitudes, influencing SC directly. Present analysis shows that the SEIO and ECP SST anomalies can contribute to SC summer rainfall variability both independently and in concert. In comparison, there are more cases of concerted contributions due to the co-variability between the Indian and Pacific Ocean SSTs.
    Wu Z. W., H. Lin, Y. Li, and Y. M. Tang, 2013: Seasonal prediction of killing-frost frequency in South-Central Canada during the cool/overwintering-crop growing season. Journal of Applied Meteorology and Climatology, 52, 102- 113.10.1175/JAMC-D-12-059.125143454-c5b1-4184-b29c-b08b86c3f7fcWOS:000313559900008a01515a1ce6643f0c3148dc4328bfc19http%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F2013JApMC..52..102Whttp://adsabs.harvard.edu/abs/2013JApMC..52..102WSeasonal killing-frost frequency (KFF) during the cool/overwintering-crop growing season is important for the Canadian agricultural sector to prepare and respond to such extreme agrometeorological events. On the basis of observed daily surface air temperature across Canada for 1957-2007, this study found that more than 86% of the total killing-frost events occur in April-May and exhibit consistent variability over south-central Canada, the country's major agricultural region. To quantify the KFF year-to-year variations, a simple index is defined as the mean KFF of the 187 temperature stations in south-central Canada. The KFF variability is basically dominated by two components: the decadal component with a peak periodicity around 11 yr and the interannual component of 2.5-3.8 yr. A statistical method called partial least squares (PLS) regression is utilized to uncover principal sea surface temperature (SST) modes in the winter preceding the KFF anomalies. It is found that most of the leading SST modes resemble patterns of El Nino-Southern Oscillation (ENSO) and/or the Pacific decadal oscillation (PDO). This indicates that ENSO and the PDO might be two dominant factors for the KFF variability. From a 41-yr training period (1957-97), a PLS seasonal prediction model is established, and 1-month-lead real-time forecasts are performed for the validation period of 1998-2007. A promising skill level is obtained. For the KFF variability, the prediction skill of the PLS model is comparable to or even better than the newly developed Canadian Seasonal to Interannual Prediction System (CanSIPS), which is a state-of-the-art global coupled dynamical system.
    Zhang H. Q., J. Qin, and Y. Li, 2011: Climatic background of cold and wet winter in southern China: part I observational analysis. Climate Dyn., 37, 2335- 2354.10.1007/s00382-011-1022-44c51c56beb3369ed0d5663975bb92c8dhttp%3A%2F%2Flink.springer.com%2F10.1007%2Fs00382-011-1022-4http://link.springer.com/10.1007/s00382-011-1022-4This study explores the climate background of anomalous wet and cold winter in southern China, focusing on results in January when most of its disastrous snowstorms and freezing rainfall events were observed. Based on the ERA-40 reanalysis and Climate Research Unit (CRU) observed precipitation and surface temperature monthly data for the period of 1959-2001, the difference between normalised monthly precipitation and temperature is used to define a simple index which reflects the intensity of the wet and cold condition in the region. It offers a good agreement with an index defined by daily weather station data observed in the region. Then, through simple correlation analyses we focus on exploring the dominant physical and dynamical processes leading to such climatic anomalies. While we acknowledge the contribution of the cold/dry air penetrated from the north, the importance of maintaining a warm and moist airflow from the south is highlighted, including an enhanced Middle East Jet Stream (MEJS) and southwesterly flow over Indochina Peninsula and South China Sea region. Strong vertical share of meridional wind, with enhanced northerly flow near the surface and southerly flow in the low to middle troposphere, leads to significant temperature and moisture inversions. These are consistent with results from synoptic analyses of the severe January 2008 event which was not included in the correlation calculations and thus suggest the 2008 event was not an unusual event although it was very intense. In the third part, we use a partial least-square statistical method to uncover dominant SST patterns corresponding to such climatic conditions. By comparing results for the periods of 1949-1978 and 1978-2007, we demonstrate the shift of dominant SST patterns responsible for the wet and cold anomalies. Shifting from "conventional" ENSO SST patterns to ENSO Modoki-like conditions in recent decades partially explains the unstable relationship between ENSO and Asian winter monsoon. Meanwhile, the importance of SST conditions in extra-tropic Pacific and Indian oceans is acknowledged. Finally, we developed a forecasting model which uses SST condition in October to predict the occurrence of the anomalous wet and cold January in the region and reasonable forecasting skill is obtained.
    Zhao P., Z. J. Zhou, and J. P. Liu, 2007: Variability of Tibetan spring snow and its associations with the hemispheric extratropical circulation and East Asian summer monsoon rainfall: An observational investigation. J.Climate, 20, 3942- 3955.10.1175/JCLI4205.12f7bc373-d7df-48c5-ab65-d52dd477151a7bab1bd4d44241d9b15cce07c37a142fhttp%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F2007JCli...20.3942Zrefpaperuri:(7f14dd985a7ce82ddca64276aaba8446)http://adsabs.harvard.edu/abs/2007JCli...20.3942ZUsing station observations of the number of days covered by snow (SCD) and snowfall over the Tibetan Plateau (TP), the National Centers for Environmental Prediction-National Center for Atmospheric Research (NCEP-NCAR) reanalysis, and precipitation from rain gauge stations in China for the period of 1973-2001, temporal/spatial variations of SCD over the TP and its associations with the hemispheric extratropical atmospheric circulation and East Asian summer monsoon rainfall are investigated. An increase of spring (April-May) SCD over the TP is associated with decreases of local tropospheric temperature and geopotential height in the spring and early summer (June). The anomalies in the tropospheric temperature and geopotential height show a westward propagation from the TP to western Asia as a result of the westward propagation of the anomalous wave energy. These tropospheric anomalies over the TP are connected with changes in the hemispheric extratropical atmospheric circulation along the westerly jet stream that acts as a waveguide. The increase of the spring SCD is also associated with the variation of the East Asian summer monsoon rainfall. Soil moisture in May-June might act as a bridge linking the spring snow anomaly and the subsequent summer monsoon. Corresponding to the increase of SCD, there is a significant decrease of the June 500-mb geopotential height from the TP to the western North Pacific. Meanwhile, the anomalous north-easterlies extend from Japan, through the east coast of China, to central-eastern China, which weaken the East Asian summer monsoon, leading to a decrease of surface air temperature and rainfall in the Yangtze and Hwai Rivers and an increase of rainfall in southeastern China. Additionally, the spring SCD anomaly is likely due to a variation of local synchronous snowfall, rather than previous winter SCD conditions. The spring SCD is not related to previous winter El Niño/La Niña events, but is associated with the equatorial central and eastern Pacific sea surface temperature from the subsequent summer through winter. The climatic implications for this relationship are not clear.
    Zhou W., C. Li, and J. C. L. Chan, 2006: The interdecadal variations of the summer monsoon rainfall over South China. Meteor. Atmos. Phys., 93, 165- 175.10.1007/s00703-006-0184-9395b56eb4ba8cfa97ee505a9c26459e0http%3A%2F%2Flink.springer.com%2Farticle%2F10.1007%2Fs00703-006-0184-9http://link.springer.com/article/10.1007/s00703-006-0184-9This paper is to promote a further understanding of the interdecadal variations of the summer monsoon rainfall over South China (SCMR). With this focus, we will specifically aim at better understanding possible mechanism responsible for such an interdecadal variation relationship between the SCMR and El Niño/Southern Oscillation (ENSO). In many of the previous studies on precipitation, the datasets used are satellite observations or gridded reanalyzed data due to the lack of long-term reliable observations over the marginal seas of the Asian continent. Such an approach could lead to possible errors in the results. In this work, several representative stations with long-term rain-gauge observations are chosen to reduce such uncertainty. The study of the interdecadal variabilities of SCMR indicates that there is a strong linkage between SCMR and ENSO on the interdecadal variations. These results agree well with those from previous studies that the Pacific Decadal Oscillation (PDO) and ENSO are not independent of each other, the interannual and interdecadal variations of tropical Pacific Sea Surface temperatures (SSTs) could affect the interdecadal variations of the SCMR, and the incorporating information on the PDO/ENSO could improve the long-term prediction of the SCMR.
    Zhu C.-W., C.-K. Park, W.-S. Lee, and W.-T. Yun, 2008: Statistical downscaling for multi-model ensemble prediction of summer monsoon rainfall in the Asia-Pacific region using geopotential height field. Adv. Atmos. Sci.,25, 867-884, doi: 10.1007/s00376-008-0867-x.10.1007/s00376-008-0867-xe3d83b8d-6066-40bc-9583-6dafc9f81e122f35c6a66db85f74728d3bcff9a5cfeehttp%3A%2F%2Flink.springer.com%2F10.1007%2Fs00376-008-0867-xrefpaperuri:(649a8a350505c555a09bf178e41c1746)http://d.wanfangdata.com.cn/Periodical_dqkxjz-e200805016.aspx
  • [1] Zhang Jijia, Chen Xingfang, 1987: THE OPERATIONAL SEASONAL FORECASTING OF THE SUMMER RAINFALL IN CHINA, ADVANCES IN ATMOSPHERIC SCIENCES, 4, 349-362.  doi: 10.1007/BF02663605
    [2] Shi Jiuen, Zhou Qinfang, Xiang Jingtian, 1986: AN APPLICATION OF THE THRESHOLD AUTOREGRESSION PROCEDURE TO CLIMATE ANALYSIS AND FORECASTING, ADVANCES IN ATMOSPHERIC SCIENCES, 3, 134-138.  doi: 10.1007/BF02680052
    [3] Congwen ZHU, Boqi LIU, Kang XU, Ning JIANG, Kai LIU, 2021: Diversity of the Coupling Wheels in the East Asian Summer Monsoon on the Interannual Time Scale: Challenge of Summer Rainfall Forecasting in China, ADVANCES IN ATMOSPHERIC SCIENCES, 38, 546-554.  doi: 10.1007/s00376-020-0199-z
    [4] LIN Zhongda, LU Riyu, 2009: The ENSO's Effect on Eastern China Rainfall in the Following Early Summer, ADVANCES IN ATMOSPHERIC SCIENCES, 26, 333-342.  doi: 10.1007/s00376-009-0333-4
    [5] Zheng HE, Pangchi HSU, Xiangwen LIU, Tongwen WU, Yingxia GAO, 2019: Factors Limiting the Forecast Skill of the Boreal Summer Intraseasonal Oscillation in a Subseasonal-to-Seasonal Model, ADVANCES IN ATMOSPHERIC SCIENCES, 36, 104-118.  doi: 10.1007/s00376-018-7242-3
    [6] DONG Haiping, ZHAO Sixiong, ZENG Qingcun, 2007: A Study of Influencing Systems and Moisture Budget in a Heavy Rainfall in Low Latitude Plateau in China during Early Summer, ADVANCES IN ATMOSPHERIC SCIENCES, 24, 485-502.  doi: 10.1007/s00376-007-0485-z
    [7] HUANG Yanyan, XUE Jishan, WAN Qilin, CHEN Zitong, DING Weiyu, ZHANG Chengzhong, 2013: Improvement of the Surface Pressure Operator in GRAPES and Its Application in Precipitation Forecasting in South China, ADVANCES IN ATMOSPHERIC SCIENCES, 30, 354-366.  doi: 10.1007/s00376-012-1270-1
    [8] TAN Jiqing, XIE Zhenghui, JI Liren, 2003: A New Way to Predict Forecast Skill, ADVANCES IN ATMOSPHERIC SCIENCES, 20, 837-841.  doi: 10.1007/BF02915409
    [9] LIN Zhenshan, SHI Xiangsheng, 2003: The Decade-Scale Climatic Forecasting in China, ADVANCES IN ATMOSPHERIC SCIENCES, 20, 604-611.  doi: 10.1007/BF02915503
    [10] CHEN Jiepeng, WU Renguang, WEN Zhiping, 2012: Contribution of South China Sea Tropical Cyclones to an Increase in Southern China Summer Rainfall Around 1993, ADVANCES IN ATMOSPHERIC SCIENCES, 29, 585-598.  doi: 10.1007/s00376-011-1181-6
    [11] John ABBOT, Jennifer MAROHASY, 2012: Application of Artificial Neural Networks to Rainfall Forecasting in Queensland, Australia, ADVANCES IN ATMOSPHERIC SCIENCES, 29, 717-730.  doi: 10.1007/s00376-012-1259-9
    [12] CHEN Lianshou, LI Ying, CHENG Zhengquan, 2010: An Overview of Research and Forecasting on Rainfall Associated with Landfalling Tropical Cyclones, ADVANCES IN ATMOSPHERIC SCIENCES, 27, 967-976.  doi: 10.1007/s00376-010-8171-y
    [13] Cai Zeyi, Wang Zuoshu, Pan Zaitao, 1992: A Numerical Study on Forecasting the Henan Extraordinarily Heavy Rainfall Event in August 1975, ADVANCES IN ATMOSPHERIC SCIENCES, 9, 53-62.  doi: 10.1007/BF02656930
    [14] Yan Shaojin, Peng Yongqing, Guo guang, 1995: Neuroid BP-type Model Applied to the Study of Monthly Rainfall Forecasting, ADVANCES IN ATMOSPHERIC SCIENCES, 12, 335-342.  doi: 10.1007/BF02656982
    [15] SUN Jianqi, Joong Bae AHN, 2011: A GCM-Based Forecasting Model for the Landfall of Tropical Cyclones in China, ADVANCES IN ATMOSPHERIC SCIENCES, 28, 1049-1055.  doi: 10.1007/s00376-011-0122-8
    [16] LOU Xiaofeng, HU Zhijin, SHI Yueqin, WANG Pengyun, ZHOU Xiuji, 2003: Numerical Simulations of a Heavy Rainfall Case in South China, ADVANCES IN ATMOSPHERIC SCIENCES, 20, 128-138.  doi: 10.1007/BF03342057
    [17] Zhina JIANG, Da-Lin ZHANG, Hongbo LIU, 2020: Roles of Synoptic to Quasi-Monthly Disturbances in Generating Two Pre-Summer Heavy Rainfall Episodes over South China, ADVANCES IN ATMOSPHERIC SCIENCES, 37, 211-228.  doi: 10.1007/s00376-019-8156-4
    [18] Haoxin ZHANG, Weiping LI, Weijing LI, 2019: Influence of Late Springtime Surface Sensible Heat Flux Anomalies over the Tibetan and Iranian Plateaus on the Location of the South Asian High in Early Summer, ADVANCES IN ATMOSPHERIC SCIENCES, 36, 93-103.  doi: 10.1007/s00376-018-7296-2
    [19] Li Chongyin, Wu Jingbo, 2000: On the Onset of the South China Sea Summer Monsoon in 1998, ADVANCES IN ATMOSPHERIC SCIENCES, 17, 193-204.  doi: 10.1007/s00376-000-0003-z
    [20] Gill M. MARTIN, Amulya CHEVUTURI, Ruth E. COMER, Nick J. DUNSTONE, Adam A. SCAIFE, Daquan ZHANG, 2019: Predictability of South China Sea Summer Monsoon Onset, ADVANCES IN ATMOSPHERIC SCIENCES, 36, 253-260.  doi: 10.1007/s00376-018-8100-z

Get Citation+

Export:  

Share Article

Manuscript History

Manuscript received: 03 December 2015
Manuscript revised: 04 April 2016
Manuscript accepted: 04 May 2016
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

A Timescale Decomposed Threshold Regression Downscaling Approach to Forecasting South China Early Summer Rainfall

  • 1. Institute of Urban Meteorology, China Meteorological Administration, Beijing 100089
  • 2. State Key Laboratory of Numerical Modeling for Atmospheric Sciences and Geophysical Fluid Dynamics, Institute of Atmospheric Physics, Chinese Academy of Sciences, Beijing 100029
  • 3. Business Intelligence & Data Analytics, Western Power, Perth WA6000, Australia

Abstract: A timescale decomposed threshold regression (TSDTR) downscaling approach to forecasting South China early summer rainfall (SCESR) is described by using long-term observed station rainfall data and NOAA ERSST data. It makes use of two distinct regression downscaling models corresponding to the interannual and interdecadal rainfall variability of SCESR. The two models are developed based on the partial least squares (PLS) regression technique, linking SCESR to SST modes in preceding months on both interannual and interdecadal timescales. Specifically, using the datasets in the calibration period 1915-84, the variability of SCESR and SST are decomposed into interannual and interdecadal components. On the interannual timescale, a threshold PLS regression model is fitted to interannual components of SCESR and March SST patterns by taking account of the modulation of negative and positive phases of the Pacific Decadal Oscillation (PDO). On the interdecadal timescale, a standard PLS regression model is fitted to the relationship between SCESR and preceding November SST patterns. The total rainfall prediction is obtained by the sum of the outputs from both the interannual and interdecadal models. Results show that the TSDTR downscaling approach achieves reasonable skill in predicting the observed rainfall in the validation period 1985-2006, compared to other simpler approaches. This study suggests that the TSDTR approach, considering different interannual SCESR-SST relationships under the modulation of PDO phases, as well as the interdecadal variability of SCESR associated with SST patterns, may provide a new perspective to improve climate predictions.

1. Introduction
  • An important task for flood and drought management is the provision of accurate rainfall prediction. The rainfall distribution in China is generally characterized by a "southern flood and northern drought" pattern, due to the weakening of the East Asian summer monsoon after the late 1970s (Nitta and Hu, 1996; Wang, 2001; Gong and Ho, 2002). Accompanied by the northward seasonal march of the East Asia summer monsoon, abundant rainfall first appears over South China (SC) in mid-May (Tao, 1987; Lau and Weng, 2001; Ding et al., 2008; Wu et al., 2012), and leads to the peak of annual rainfall in early summer (June) over the region (Fig. 1a). Because SC is one of China's largest economic zones with a large population, extreme flood events in early summer often cause a large number of human casualties and

    considerable economic loss (Chan and Zhou, 2005; Zhou et al., 2006). Thus, improving seasonal forecasting skill for early summer rainfall over SC is of great importance for disaster prevention and mitigation.

    Figure 1.  (a) The annual climatological cycle of the rainfall at the stations of Guangzhou, Hong Kong and Macao, as well as the 3-station average, during 1910-2011. (b) Detrended correlation of the 3-station averaged rainfall with the 160-station rainfall from CMA in June during 1951-2011. The stations at Guangzhou, Hong Kong and Macao are indicated by the three red dots. The grey shading indicates CCs significant at the 99% confidence level based on the t-test, presenting the SC region. (c) Power spectrum for June rainfall. The peak over the red dashed line indicates the confidence level is greater than 80% against red noise.

    There are several approaches to forecasting rainfall. One is using numerical models (e.g. Collischonn et al., 2005; Aligo et al., 2009). The raw model prediction is dynamically meaningful, but still has low skill for many reasons; for instance, because of the model resolution (Martin, 1999), subgrid processes (Grotch and MacCracken, 1991) and parameterization schemes (Eitzen and Randall, 1999). Hence, precipitation modeling is regarded as one of the most difficult issues in climate modeling, resulting in low skill on the basis of the pure numerical approach. Another approach to forecasting rainfall is based on statistical downscaling models (Liu and Fan, 2012a, 2014; Sun and Chen, 2012). A statistical downscaling method combining the GCM-simulated and observed information is developed, which shows much better predictability for global precipitation forecasting (Sun and Chen, 2012). In addition, pure statistical downscaling models are also used to forecast rainfall. The method is generally based on an empirical observed relationship between the large-scale climate anomalies and local rainfall fluctuations. There are various methods that can be used to develop statistical downscaling models, including multiple linear regression (Wilby, 1998), principle components (Li and Smith, 2009), and singular value decomposition (Zhu et al., 2008; Liu and Fan, 2012b). Other more sophisticated methods include partial least squares (PLS) regression (Zhang et al., 2011; Wu et al., 2013).

    Pure statistical downscaling models show reasonable skill in predicting regional rainfall. For example, (Sahai et al., 2003) made optimum use of global SST for Indian summer monsoon rainfall prediction nine months in advance. Rainfall may contain variabilities on various timescales, with low-frequency interdecadal variability associated with pronounced wetting or drying trends (Ding et al., 2008) and high-frequency interannual variability related to severe floods or droughts (Huang et al., 2006). (Guo et al., 2012) described a timescale decomposition (TSD) approach to statistically downscale the late summer rainfall over North China, which made use of two distinct downscaling models corresponding to the interannual and interdecadal rainfall variability, respectively. However, the TSD approach by (Guo et al., 2012) neglected the possibility that the interannual rainfall variability may be modulated by an interdecadal climatic background.

    Recently, an increasing number of studies have suggested that the interannual relationships of large-scale climate anomalies with local or remote climate fluctuations are not stationary (e.g. Torrence and Webster, 1999; Wang, 2002; Wu and Wang, 2002; Gao et al., 2006; Sun and Wang, 2012; Chen et al., 2013a; Chen et al., 2015a, 2015b; Cao et al., 2015, 2016). In particular, the modulation of a decadal-scale coupled ocean atmospheric mode named the Pacific Decadal Oscillation (PDO) (Mantua et al., 1997; Mantua and Hare, 2002) has attracted more attention in the last decade (Gershunov and Barnett, 1998; Power et al., 1999; Chan and Zhou, 2005; Wang et al., 2008; Mao et al., 2011; Chen et al., 2013b; Duan et al., 2013). For example, it has been found that the typical influences of ENSO on the North American climate are strong and consistent only during preferred phases of the PDO (Gershunov and Barnett, 1998). The ENSO-East Asian summer monsoon relationship during 1962-77 has been found to be significantly different from that during 1978-93 (Wu and Wang, 2002), which is actually consistent with a phase transition from negative to positive PDO in the late-1970s. In the SC region, the interannual rainfall variation has also been suggested to be modulated by the PDO (Chan and Zhou, 2005; Mao et al., 2011; Duan et al., 2013). It has been proposed that incorporating information on the PDO could improve the long-term predictability of early summer SC rainfall changes (Chan and Zhou, 2005; Zhou et al., 2006). Therefore, a statistical model that takes into account the modulation of the PDO may lead to better skill in seasonal rainfall prediction over SC. This naturally raises two key questions: How can the modulation effect of the decadal-scale coupled oceanic-atmospheric mode of the PDO be incorporated when developing a statistical downscaling model to forecast the early summer rainfall over SC? And can the forecasting skill be improved?

    The aim of this paper is to describe a TSD threshold regression (TSDTR) downscaling approach for the statistical forecasting of South China early summer rainfall (SCESR). The TSDTR downscaling approach includes two distinct regression downscaling models, called the interannual model (IAM) and interdecadal model (IDM), respectively linking SST patterns prior to early summer to the interannual and interdecadal variability of SCESR. On the interannual timescale, the IAM, based on a threshold PLS regression model, is fitted to interannual components of SCESR and March SST patterns by taking account of the modulation of negative and positive phases of the PDO. On the interdecadal timescale, the IDM, based on a PLS regression model, is employed to fit the relationship between SCESR and preceding November SST patterns. The total rainfall prediction is obtained by the sum of the outputs from both the IAM and IDM.

    The rest of this paper is arranged as follows: Section 2 introduces the datasets used in this work. The TSDTR downscaling approach is described in section 3. Results of the TSDTR approach in forecasting SCESR are presented in section 4. Section 5 is a summary and discussion.

2. Data
  • Long-term reliable observed rainfall data are derived from the monthly rain gauge datasets from three stations: Hong Kong (22.11°N, 114.14°E), Macao (22.16°N, 113.35°E) and Guangzhou (23.08°N, 113.16°E). These rainfall data have been used by previous studies (Chan and Zhou, 2005; Mao et al., 2011; Duan et al., 2013) and are chosen here because of their long-term reliability since 1910. The maximum rainfall of these three stations occurs in June (Fig. 1a), which purely belongs to the South China Sea summer monsoon rainfall (Tao, 1987; Wang et al., 2004). Another set of monthly rainfall data derived from 160 Chinese meteorological stations provided by the China Meteorological Administration (CMA) is also used in this study, available from 1951. The simultaneous correlation pattern of three-station averaged rainfall in June with the 160-station rainfall from CMA during the period 1951-2011 shows significant positive correlation over the region of SC (Fig. 1b). Thus, it is reasonable to use the long-term, high-quality three-station (Hong Kong, Macao and Guangzhou) averaged rainfall in June to represent the SCESR in this study.

    The PDO index, defined as the leading EOF of SST anomalies in the North Pacific Ocean poleward of 20°N after removing the global warming signal (Mantua et al., 1997; Mantua and Hare, 2002), is taken from the website of the Joint Institute for the Study of the Atmosphere and Ocean (http://jisao.washington.edu/pdo/PDO.latest), which contains data from 1900 onward. An 11-year running mean of wintertime (October-March averaged) PDO index is used to obtain the interdecadal variability of the PDO (Mantua and Hare, 2002; Mao et al., 2011) over the period 1910-2011 (Fig. 2a). Note that 10 years are missing, caused by the 11-year running mean, and thus the period for the forecasting experiment in this work is chosen as 1915-2006. Monthly SST data are extracted from NOAA ERSST.v3b (Smith and Reynolds, 2004; Smith et al., 2008) on 2.0°× 2.0° grid, which is available from the year 1854 at http://www.esrl.noaa.gov/psd/data/gridded/.Monthly atmospheric data are obtained for the period from 1948 from the NCEP-NCAR reanalysis products on 2.5°× 2.5° grid (Kalnay et al., 1996), including horizontal winds and specific humidity at different levels (also available at http://www.esrl.noaa.gov/psd/data/gridded/).

    Figure 2.  (a) The standardized time series of wintertime (October-March) PDO index and their 11-year running mean during 1915-2006. As an example, the year 1915 denotes October-December 1914 and January-March 1915. (b) Total rainfall (Rain$_\rm T$, units: mm) in June during 1915-2006, which is decomposed into an interannual component (Rain$_\rm A$) with variation less than 11 years and an interdecadal component (Rain$_\rm D$) with variation longer than 11 years.

3. Methods
  • Two primary peaks with periods of about 4 years and 17 years exist in the SCESR, as determined by spectrum analysis (Fig. 1c), indicating apparent interannual and interdecadal variability. Therefore, the observed time series of total rainfall (Rain T) is decomposed into interannual (variation less than 11 years, denoted as Rain A) and interdecadal (variation longer than 11 years, denoted as Rain D) components by an 11-year high-pass and low-pass filter (Fig. 2b). That is, Rain T= Rain A+ Rain D.

    SST anomalies have been found to be important in influencing the East Asian climate (Huang and Wu, 1989; Wang et al., 2000; Lau and Weng, 2001; Wu et al., 2012), have a relatively long "memory" (Sahai et al., 2003), and can therefore be regarded as a preceding predictor for SCESR. This motivates us to explore the relationships between SCESR and associated SST patterns on both interannual and interdecadal timescales. As such, we also decompose the SST field into interannual (SST A) and interdecadal (SST D) components. To reveal the dominant SST patterns associated with SCESR variations, we employ the PLS regression method. Specifically, PLS embodies the well-known concept of partial correlation, as it seeks the predictors Z, which are linear combinations of the factors X, being referred to as latent vectors or PLS components, and maximizes the variance explained in Y and the correlation between X and Y (Haenlein and Kaplan, 2004; Smoliak et al., 2010). Unlike EOF analysis, which identifies major patterns explaining SST variations, using PLS regression we can find the PLS components of SST variations that best explain the covariance between SST variations and the SCESR variations. In other words, it reveals the dominant SST patterns that not only account for most of the SST variations but are also closely related to the SCESR variability.

    Figure 3.  Schematic plot of the TSDTR approach to forecasting the SCESR.

    Figure 3 shows a schematic plot of the TSDTR approach to forecasting the SCESR based on the PLS regression method. On the interannual timescale, in order to incorporate the modulation effect of the PDO on the interannual variability of SCESR (Mao et al., 2011; Duan et al., 2013), a threshold PLS regression model is calibrated for the relationship between Rain A and SST A under the positive and negative phase of the PDO. On the interdecadal timescale, a standard PLS regression model is calibrated for the relationship between Rain D and associated SST D patterns. The total rainfall prediction is obtained by the sum of the outputs \(\hat{R}ain_\rm A\)(t) and \(\hat{R}ain_\rm D\)(t) from both the IAM and IDM (Fig. 3). To test the performance of the TSDTR approach in forecasting the SCESR, the study period 1915-2006 (N=92) is separated into a calibration period [1915-84 (n=70)] and validation period (1985-2006). The two PLS-based regression downscaling models in the TSDTR approach are calibrated by using the calibration data in 1915-84, and the forecasting skill of the TSDTR is tested by the independent validation data in 1985-2006. To obtain the forecasted values in the independent validation period, we use the running forecasting method based on the calibrated models. That is, for t=n+1,n+2,…,N (=92), when the observed preceding SST(t) data and wintertime PDO are available over the validation period 1985-2006, we add these new SST data to those in the training period, and then decompose the combined SST data into interannual SST A(t) and interdecadal SST D(t) components. The forecasted value of interannual [interdecadal] rainfall can be estimated by using SST A(t) [SST D(t)] and the calibrated IAM [IDM]. The performance of the TSDTR downscaling approach is assessed through the correlation coefficient (CC) between predicted and observed values and the RMSE (Zhang et al., 2011; Guo et al., 2012). The uncertainty of the forecast is indicated by the spread of bootstrapping prediction intervals [see appendix in (Li and Smith, 2009)].

4. Forecasting SCESR
  • This section provides details of the TSDTR downscaling approach to building the IAM and IDM based on the PLS regression for the relationship between the preceding month SST and the SCESR on the interannual and interdecadal timescales, respectively.

  • The interannual variation of SCESR is remarkably different under different PDO phases. (Chan and Zhou, 2005) demonstrated that the interannual relationship between ENSO and SCESR is modulated by the phase of the PDO. (Mao et al., 2011) indicated that the interannual SCESR variations are remarkably different under different PDO phases, based on a comparison in two typical epochs of 1958-76 and 1980-98. Furthermore, it was suggested by (Duan et al., 2013) that the predictability of interannual SCESR variability is modulated by different PDO-phase backgrounds, and they also indicated that the relationship between interannual rainfall and preceding Pacific SST anomalies experienced a robust interdecadal change due to the PDO's modulation through the so-called "seasonal footprinting mechanism".

    Figure 4.  Lead-lag correlation of Rain$_\rm A$ with SST$_\rm A$ in the Pacific and Indian oceans in the (a-c) negative phase and (d-f) positive phase of the PDO. The contour lines present the CCs of $\pm0.3$, statistically significant at the 95% confidence level.

    Figure 5.  First two leading modes of March SST$_\rm A$ in the Pacific and Indian oceans from PLS regression analyses in the (a, b) negative phase and (c, d) positive phase of the PDO. The first percentage value is the total variance of Rain$_\rm A$ explained by the SST$_\rm A$ mode, and the second value is the total SST$_\rm A$ variance explained by the same mode.

    To further explore the predictability of interannual SCESR variability modulated by different PDO-phase backgrounds, Fig. 4 shows the lag-correlation patterns of interannual SCESR with preceding interannual monthly (May, April and March) SST A over the Pacific and Indian oceans in negative and positive phases of the PDO, denoted by PDO(-) and PDO(+), respectively. It is evident that the interannual relationship between Rain A and SST A exhibits very different structures during the positive and negative PDO phases. A striking difference in Fig. 4 is that there is a traditional eastern Pacific warming ENSO-like correlation pattern between Rain A and SST A during the PDO(+) phase, but a central Pacific warming ENSO-like pattern is more pronounced during the PDO(-) phase. Such correlation patterns may lead the SCESR by up to 3 months from March to May. This result is consistent with the result of (Duan et al., 2013), who compared 1955-76 PDO(-) and 1977-98 PDO(+). The result supports the fact that the interannual SCESR variability and its relationship with ENSO are modulated by the PDO phases (Chan and Zhou, 2005; Zhou et al., 2006; Mao et al., 2011). Therefore, the interannual relationship between large-scale SST anomalies and the SCESR is not stationary due to the modulation of the PDO. As such, in order to incorporate the modulation effect of the PDO on the interannual variability of SCESR, we need to consider the nonstationary relationship between Rain A and SST A in the Pacific and Indian oceans. To this end, a threshold PLS regression model (Fig. 3) is employed to establish the relationship between Rain A and the leading SST A modes, which best explain the co-variations of Rain A and SST A under the PDO(-) and PDO(+), separately. After a series of tests using SST A predictors in each preceding month, including the preceding May, April, March, February, January, December and November, it is found that SST A conditions in the preceding March tend to yield the best results. Thus, we concentrate on reporting the dominant SST A modes in March over the Pacific and Indian oceans that influence Rain A.

    Figure 5 shows the first two leading interannual modes of March SST (SST A-I and SST A-II) represented by PLS loadings in PDO(-) and PDO(+), respectively. The significance of each mode is reflected by two numbers: the first is the percentage of Rain A variance explained by the SST-PLS mode; the other is the percentage of SST A variance explained by the same SST-PLS mode. In PDO(-), the first two leading modes together explain 73.2% of the total variance of Rain A. The first dominant mode (Fig. 5a) during the negative phase of the PDO is, by and large, an El Niño Modoki-like pattern in the tropical Pacific region, with positive loadings in the central Pacific and negative loadings in the western and eastern Pacific. Another significant positive loading occurs over the South-central Pacific region around New Zealand. The tropical Indian Ocean is characterized by weak positive loadings. As shown in Fig. 6a with respect to the corresponding patterns of precipitation and vertically integrated moisture flux to the SST A-I mode, notable southwesterly moisture transportation to the SC region is observed in the troposphere, which is accompanied by a significant low-level anomalous anticyclone located in the South China Sea and Philippine Sea (not shown). At the same time, southward penetration of northerly moisture transportation from a large part of North and Northeast China also exists and leads to significant moisture convergence over the SC region. Therefore, abundant interannual rainfall is received in this region (Fig. 6a).

    The second dominant SST mode (Fig. 5b) in PDO(-) has strong negative loadings in the central-eastern tropical Pacific and weaker negative loadings in the tropical Indian Ocean, indicating a La Niña-like pattern. Signals can also be found in the midlatitudes, with anomalous warming in the North Pacific region and South-central Pacific region around New Zealand. The pattern in the tropical and midlatitude North Pacific resembles the anomalous SST features of negative PDO phase. Correspondingly, one can see notable southerly moisture transportation over the majority of eastern China (Fig. 6b), which is accompanied by an anomalous low-level anticyclone located in the north to northeast of the Philippine Sea (not shown). The SC region is under the control of southerly moisture transport, but the moist flows are strong enough to advance more northward. Thus, distribution of anomalous precipitation in China shows weak above-normal interannual rainfall over SC (Fig. 6b).

    Figure 6.  Correlations of the detrended scores of March SST$_\rm A$ modes in Fig. 5 with high-pass filtered June precipitation (color shading) and vertically integrated moisture flux between 1000 hPa and 300 hPa (vectors) in the (a, b) negative phase and (c, d) positive phase of the PDO.

    Figure 7.  (a) PLS regression model results for forecasting Rain$_\rm A$. PDO negative (positive) years are marked with "o" ("*"). The black line is the time series of observed Rain$_\rm A$; the red line is the reconstructed Rain$_\rm A$ during 1915-84; and the blue line is the hindcasted Rain$_\rm A$ during 1985-2006. (b) PLS regression model results for forecasting Rain$_\rm D$. The black line is the time series of observed Rain$_\rm D$; the red line is the reconstructed Rain$_\rm D$ during 1915-84; and the blue line is the hindcasted Rain$_\rm D$ during 1985-2006. (c) Model results for forecasting the total rainfall. The black line is the time series of observed Rain$_\rm T$; the red line is the reconstructed Rain$_\rm T$ during 1915-84; and the blue line is the hindcasted Rain$_\rm T$ during 1985-2006. The blue shading in each of the panels represents the upper and lower bands of the bootstrapping 95% confidence intervals for the hindcasted values.

    In PDO(+), the first two leading modes together explain 58.3% of the total variance of interannual rainfall. The leading interannual SST modes (SST A-I and SST A-II) dramatically change in contrast with those in PDO(-). The first dominant mode (Fig. 5c) tends to suggest a strengthened traditional El Niño-like signal in influencing the interannual early summer rainfall over SC during the positive phase of the PDO, with very strong positive loadings in the tropical eastern Pacific. Another positive loading occurs in the tropical Indian Ocean. In addition, the positive loading in the South-central Pacific region around New Zealand in PDO(-) is replaced by negative loading. Meanwhile, the pattern in the midlatitude North Pacific shares similar features with SST anomalies during the positive phase of the PDO. The circulation responses to this mode (Fig. 6c) show notable southwesterly moisture transportation from the South China Sea to the southeast coast of China, which leads to above-normal precipitation over the SC region. The second dominant interannual SST mode (SST A-II) in PDO(+) has negative loadings in the central-eastern tropical Pacific that display a La Niña-like pattern (Fig. 5d), but has reduced in this mode compared with that in PDO(-) (Fig. 5b). The Indian Ocean is occupied by weak negative loadings. In addition, a warming anomaly occurs in Peru's inshore waters. Figure 6d shows the corresponding interannual circulation responses. The vertically integrated moisture flux pattern indicates that the main moisture source region for interannual rainfall over SC in this mode is from the Bay of Bengal. Through the westerly flows carrying warm and wet air, the SC region tends to receive above-normal interannual rainfall.

    Based on the above analysis, the different leading interannual SST modes related to the interannual early summer rainfall over SC in PDO(-) and PDO(+) may indicate that different PDO "backgrounds" modulate the connection between the interannual early summer SC rainfall and SST anomalies over the Pacific and Indian oceans. Therefore, when building a statistical forecasting model for interannual rainfall with preceding SST conditions, one needs to consider such nonlinearity. As previously mentioned, a threshold PLS regression model (Fig. 3) is employed to establish the relationship between Rain A and the leading SST A modes by taking into account the modulation of PDO(-) and PDO(+).

    Figure 7a compares the interannual variation of predicted and observed Rain A values from the threshold PLS regression model (i.e. the IAM in Fig. 3) based on the first two leading interannual SST modes, which is derived using observed data over the training period of 1915-84 (n=70). The results indicate that Rain A can be reconstructed with significant skill for the training period (Fig. 7a), with a CC0 of 0.87 and RMSE0 of 54.45 mm between the predicted and observed Rain A values (Model IV in Table 1). Here, CC0 and RMSE0 are the CC and RMSE between the observed rainfall and reconstructed rainfall in the calibration period. To obtain the forecasted values [\(\hat{R}ain_\rm A\)(t)] in the independent validation period of 1985-2006, we use the running forecasting method based on the calibrated IAM. The forecasted value of interannual rainfall [\(\hat{R}ain_\rm A\)(t)] can be estimated by using SST A(t) and the calibrated IAM (i.e. the calibrated threshold PLS regression model based on the preceding wintertime PDO phases). Figure 7a shows the forecasted values [\(\hat{R}ain_\rm A\)(t)] in the validation period of 1985-2006 and their uncertainty in terms of bootstrapping 95% confidence intervals.

    Figure 8.  First two leading modes of November SST$_\rm D$ in the western-central North Pacific from PLS regression analyses. The first percentage value is the total variance of Rain$_\rm D$ explained by the SST$_\rm D$ mode and the second number is the total SST$_\rm D$ variance explained by the same mode.

  • Next we develop an IDM for the interdecadal variability of the SCESR. Previous studies have noted that temporal variations of the SCESR exhibit an interdecadal oscillation related to the PDO, with more dry (wet) years during periods of positive (negative) PDO index (Chan and Zhou, 2005; Duan et al., 2013). (Chan and Zhou, 2005) also mentioned that the effect of the PDO is more important than that of ENSO in the control of SCESR, although such a conclusion might be premature given our limited understanding of the PDO and ENSO, as well as their possible interaction. Because the rainfall and PDO index have a close negative relationship on the interdecadal timescale, and the most visible climatic fingerprints of the PDO exist in the North Pacific (Mantua et al., 1997; Mantua and Hare, 2002), it is supposed that the most important interdecadal SST signal influencing the interdecadal rainfall variability may exist in the midlatitude North Pacific region. With the aim to develop a PLS-based regression downscaling model with preceding interdecadal SST conditions to forecast the interdecadal rainfall variability, a series of tests using SST A predictors in each preceding month, including the preceding May, April, March, February, January, December and November, are conducted, and the results show that interdecadal SST conditions over the midlatitude western-central North Pacific [(20°-50°N, 120°E-140°W)] in the preceding November give the best results. Thus, we focus on reporting the leading interdecadal SST modes in November over the western-central Pacific region and the corresponding atmospheric circulation responses to the SST modes influencing the interdecadal SC rainfall variability. To this end, the PLS regression model (Fig. 3) is employed to establish the relationship between Rain D and the leading SST D modes using the training data in the period 1915-84.

    Figure 8 shows the first two leading interdecadal SST modes (SST D-I and SST D-II) represented by PLS loadings. These two leading modes together explain 74.1% of the variance of Rain D, with the majority contributed from the first leading mode (58.1%). The SST D-I mode (Fig. 8a) is characterized by significant positive loadings that almost cover the entire western-central North Pacific, with two maximum anomalous SST centers located in the Yellow Sea and midlatitude central North Pacific located at about (35°N, 160°W). However, these positive loadings in the western-central North Pacific shrink in the second mode (SST D-II); instead, negative loadings widely expand (Fig. 8b). The SST D-I mode resembles the characteristics of the PDO negative phase to some extent, with a typical warm SST anomaly pattern in the central North Pacific (Mantua et al., 1997; Mantua and Hare, 2002). This is therefore consistent with previous results in which persistent wet conditions over the SC are related to periods of negative PDO index (Chan and Zhou, 2005; Zhou et al., 2006). In the second dominant SST mode (SST D-II), the shrinking of positive loadings may imply a phase transformation from negative to positive PDO.

    To examine the corresponding atmospheric circulation anomalies associated with the above two dominant SST modes on the interdecadal timescale, we show the correlation patterns of the scores of SST modes with the vertically integrated moisture flux between 1000 hPa and 300 hPa (Fig. 9). The anomalous precipitation distributions in China are also displayed in Fig. 9. For the SST D-I mode, remarkable southwesterly moisture flux enters into the SC region, which would lead to persistently wet conditions in the region (Fig. 9a). As a result, a significant positive correlation with interdecadal rainfall variability over the SC region ultimately occurs (Fig. 9a). The atmospheric circulation responses to the second mode (SST D-II) are basically similar to those for the first mode, only with small differences in magnitude, thus also exerting a beneficial influence on above-normal interdecadal rainfall variability over the SC region (Fig. 9b). The SST anomalies in the North Pacific associated with the PDO can persist for 20-30 years, from winter to summer (Mantua et al., 1997), and thus seem to be important in the control of the persistently wet or dry conditions in early summer over SC. Note that since the anomalies associated with persistently wet and dry conditions generally tend to have opposite polarities, the reverse is true for interdecadal below-normal rainfall variability.

    Figure 7b compares the interdecadal variation of predicted and observed Rain D values from the PLS regression model (i.e. the IDM in Fig. 3) based on the first two leading modes (SST D-I and SST D-II), which is derived using observed data over the training period of 1915-84 (n=70). It is evident that the IDM can reproduce the time series of interdecadal rainfall variability in the training period well, with a CC0 of 0.88 and RMSE0 of 62.67 mm between the predicted and observed values of Rain D. To obtain the forecasted interdecadal values [\(\hat{R}ain_\rm D\)(t)] in the independent validation period of 1985-2006, we use the running forecasting method based on the calibrated IDM. The forecasted value of interdecadal rainfall [\(\hat{R}ain_\rm D\)(t)] can be estimated by using SST D(t) and the calibrated IDM. Figure 7b shows the forecasted values [\(\hat{R}ain_\rm D\)(t)] in the validation period of 1985-2006 and their uncertainty in terms of bootstrapping 95% confidence intervals.

  • It is straightforward to forecast the values of the SCESR (i.e. Rain T) by summing up the forecasted values \(\hat{R} ain_\rm A\)(t) and \(\hat{R}\rm ain_\rm D \)(t) from the IAM and IDM models (i.e. the TSDTR downscaling approach). Figure 7c shows the performance of forecasting the SCESR. In general, the performance in the training period is maintained in the subsequent validation period. For example, compared to the observed climatological rainfall of 343.80 mm during the training period of 1915-84 and 367.70 mm during the validation period of 1985-2006, the statistical model provides a preproduction of 343.80 mm and 351.15 mm, respectively. The TSDTR downscaling approach provides reasonable forecasting skill by using preceding SST as the only predictor. Table 1 shows the forecasting skill of total rainfall by showing the associated CC ( CC0=0.88) and RMSE ( RMSE0=62.67 mm) between the downscaled and observed rainfall for the training period of 1915-84. This skill is maintained reasonably well in the validation period with a CC1 of 0.56 and RMSE1 of 139.47 mm (Table 1, Model IV). Here, CC1 and RMSE1 are the CC and RMSE between the observed rainfall and predicted rainfall in the validation period. The bootstrapping 95% confidence intervals associated with the forecasted values are shown in Fig. 7c, indicating the uncertainty of the TSDTR downscaling approach. All these results indicate that the TSDTR downscaling approach can achieve some reasonable rainfall forecasting skill over SC.

    Figure 9.  Correlations of the detrended scores of November SST$_\rm D$ modes in Fig. 8 with low-pass filtered June precipitation (color shaded) and vertically integrated moisture flux between 1000 hPa and 300 hPa (vectors).

    To see how well the TSDTR downscaling approach performs in forecasting the SCESR, compared to other simple methods, we examine three different methods for forecasting the SCESR. We analyze the use of the previous 5 years' SCESR average (P5YSA), a single PLS model, and a TSD-PLS model. The P5YSA is a computationally convenient and frequently used technique to forecast a regular time series, which is carried out without recourse to a formal statistical model. The single PLS model is calibrated by directly regressing the SCESR onto the SST field, while the TSD-PLS model is calibrated by two steps; namely, decomposing the rainfall and SST field into interannual and interdecadal components, and then applying the PLS method to them. All these three methods do not integrate the modulation of the PDO on the interannual relationship between SCESR and SST. Note that the first two dominant SST patterns are used in both the PLS and TSD-PLS models for a fair comparison with the TSDTR downscaling approach. Table 1 summarizes how the forecasting skill of the TSDTR downscaling approach compares with the three simpler models. The CC between the forecasted and observed SCESR values in the validation period for P5YSA, PLS and TSD-PLS are CC1=0.13, 0.06 and 0.46, respectively, which is smaller than that (0.56) from the TSDTR downscaling approach. In addition, the RMSE between the forecasted and observed SCESR values in the validation period for P5YSA, PLS and TSD-PLS are RMSE1=167.22, 183.81 and 148.06 mm, respectively——larger than the RMSE ( RMSE1=139.47 mm) for the TSDTR downscaling approach. Therefore, the more complex TSDTR downscaling approach performs better than the simpler P5YSA, PLS and TSD-PLS models in forecasting the SCESR. We thus emphasize that the application of the newly proposed TSDTR method can improve regional rainfall prediction skill by incorporating the PDO modulation effect, even when the skill of the pure PLS model is low.

    In summary, the rationale behind the TSDTR downscaling approach is that it allows a model parameter change for the unstable relationship between the interannual rainfall variability and preceding SST conditions in different PDO phases. As a result, the TSDTR downscaling approach can further improve rainfall forecasting skill by considering the modulation effect of decadal-scale coupled oceanic-atmospheric modes on interannual climate variability, compared to other simpler methods (e.g. P5YSA, PLS and TSD-PLS), without considering the preceding SST conditions in different PDO phases.

5. Summary and discussion
  • This paper puts forward the TSDTR downscaling approach to forecast SCESR by using preceding SST over the Indian and Pacific oceans through modeling the interannual and interdecadal rainfall variability via an IAM and IDM. The IAM is built upon a threshold PLS regression by taking account of the modulation effect of the PDO on the interannual variability of the SCESR. Through the IAM, the first two leading March SST modes (an El Niño Modoki-like pattern and a La Niña-like pattern) are linked to the interannual rainfall in the negative phase of the PDO, while the first two leading modes (a strengthened traditional El Niño-like pattern and a La Niña-like pattern) are linked to the interannual variability of the SCESR in the positive phase of the PDO. The first two leading interdecadal SST modes are linked to the interdecadal component of the SCESR via the IDM based on the PLS regression method. The forecasted total rainfall is obtained by summing up the values of the two forecasted components from the IAM and IDM.

    In applying the TSDTR approach to forecasting SCESR, it is found that the interannual relationship between the preceding Pacific-Indian Ocean SST and SCESR experiences interdecadal changes, displaying remarkable differences in the negative and positive phases of the PDO. In particular, the traditional ENSO-like anomalous SST pattern is robust only in the PDO positive phase. For interdecadal rainfall variability, the dominant interdecadal SST patterns over the North Pacific resemble some characteristics of the PDO. When the North Pacific SST is persistently warming (cooling), the interdecadal variability of SCESR tends to be below-normal (above-normal). This is consistent with previous studies in which persistently wet (dry) conditions over SC in early summer are related to periods of negative (positive) PDO index (Chan and Zhou, 2005; Duan et al., 2013).

    A central implication of this study is that regional rainfall forecasting skill can be improved via the TSDTR approach, wherein the unstable relationship between interannual rainfall and large-scale variability, such as SST, can be taken into the process of model development. For a statistical downscaling model, it is important to select the predictors that have the most stable relationships with observed rainfall (Sun and Chen, 2012): If the predictor has a stable relationship with the predictand, the predictability of the statistical downscaling scheme developed using this predictor will be stable; otherwise, a predictor with an unstable relationship with the predictand could result in unstable predictability of the statistical downscaling scheme. As the TSDTR approach considers the modulation effect of decadal-scale coupled oceanic-atmospheric modes on interannual climate variability, it may provide a new perspective to improve climate prediction. In this paper, we demonstrate that the TSDTR downscaling forecasting skill is superior to three other simpler methods (P5YSA, PLS and TSD-PLS), without considering the preceding SST conditions in different PDO phases. The results suggest that the TSDTR approach has a higher predictive capability SCESR.

    In spite of the improvement in forecasting skill, the TSDTR downscaling forecasting model for SCESR encounters a problem in that it fails to predict some extreme flood events, such as in 2001. This failure is attributable to the interannual rainfall model, because the preceding interannual spring SST condition used as the predictor in 2001 displays no significant anomalies (not shown), indicating that the SST anomaly itself is not the only factor that can exert an important influence on the interannual SC rainfall. Thus, research gaps exist insofar as other potential factors need to be considered in the development of the TSDTR model. We expect in the future to include some of these other factors [e.g. the atmospheric circulation variability over the extratropics and tropics, and soil moisture conditions (e.g. Thompson and Wallace, 1998; Zhao et al., 2007; Chen et al., 2014)], with the aim to improve the statistical forecasting skill further.

  • PLS regression

    PLS regression is usually described as a two-staged approach. Following \cite{Butler2000},the first stage is to produce a sequence of $k\le m$ PLS components $Z_i$ of length $n$ to be included in the regression. However, unlike principal components, which are only formed by accounting for the maximum amount of joint variability of ${X}$, the PLS component $Z_i$($i=1,2,\ldots,k$) is chosen to explain as much as possible the covariance between ${X}$ and $Y$. That is,$$Z_i={Xc}_i \ \ (A1)$$ is chosen to maximize $Z'_iY$ subject to the loadings ${ c}_i$ that $\|{ c}_i\|=1$ and that $Z_i$ is orthogonal to the space spanned by the basis $\{Z_1,Z_2,\ldots,Z_{i-1}\}$. The second stage of PLS regression is to regress $Y$ on the PLS components $Z_i$($i=1,2,\ldots,k$), which gives a simple linear model $$ Y=\sum_{i=1}^k {b_iZ_i}+\varepsilon_k~,(A2)$$ where $\varepsilon_k$ is the appropriate error term. The parameter $b_i$ is derived by minimizing the sum of squares,$$\left(Y-\sum_{i=1}^k{b_iZ_i}\right)'\left(Y-\sum_{i=1}^k{b_iZ_i}\right)=(Y-X\beta_{{\rm PLS}})'(Y-X\beta_{{\rm PLS}})~,(A3)$$ where $\beta_{{\rm PLS}}=\sum_{i=1}^k{b_ic_i}$ is the PLS parameter vector. From \cite{Helland1988},we have $$\hat{\beta}_{{\rm PLS}}=\sum_{i=1}^k\hat{\gamma}_i({X}'{X})^{i-1}{X}'Y~,$$ where the parameter $\hat{\gamma}_i$ can be estimated by using the PLSR1 algorithm. Predictions $\hat{y}_{x'}$ of future responses can then be made by $$ \hat{y}_{x'}=\overline{y}+\sum_{j-1}^k\hat{\beta}_{{\rm PLS},j}(x'_j-\overline{x}_j)~.$$

    Note that the selected $Z_i$($i=1,2,\ldots,k$) is obtained by accounting for the maximum amount of the covariance between ${X}$ and $Y$. Thus, PLS components are obtained by not only accounting for the variance of explanatory variables ${X}$, but also the variances in the predicant $Y$.

Reference

Catalog

    /

    DownLoad:  Full-Size Img  PowerPoint
    Return
    Return