Advanced Search
Article Contents

Impacts of Two Types of El Niño on the MJO during Boreal Winter


doi: 10.1007/s00376-016-5272-2

  • The features of the MJO during two types of El Niño events are investigated in this paper using the daily NCEP-2 reanalysis data, OLR data from NOAA, and Real-time Multivariate MJO index for the period 1979-2012. The results indicate that the MJO exhibits distinct features during eastern Pacific (EP) El Niño events, as compared to central Pacific (CP) El Niño events. First, the intensity of the MJO is weakened during EP El Niño winters from the tropical eastern Indian Ocean to the western Pacific, but enhanced during CP El Niño winters. Second, the range of the MJO eastward propagation is different during the two types of El Niño events. During EP El Niño winters, the MJO propagates eastwards to 120°W, but only to 180° during CP El Niño winters. Finally, the frequency in eight phases of the MJO may be affected by the two types of El Niño. Phases 2 and 3 display a stronger MJO frequency during EP El Niño winters, but phases 4 and 5 during CP El Niño winters.
  • 加载中
  • Ashok K., S. K. Behera, S. A. Rao, H. Y. Weng, and T. Yamagata, 2007: El Niño Modoki and its possible teleconnection. J. Geophys. Res.: Oceans, 112, C11007.10.1029/2006JC00379840d0c13a-a855-452f-b9dc-0947592e2d099ed3e543c8e9bfe16c01f5a1395c1babhttp%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1029%2F2006JC003798%2Fabstract%3Bjsessionid%3DF2C5797F7539FC367A07E919851C7401.f03t02refpaperuri:(42a43d3a8b5a3852d00d87760b9d78f8)http://onlinelibrary.wiley.com/doi/10.1029/2006JC003798/abstract;jsessionid=F2C5797F7539FC367A07E919851C7401.f03t02[1] Using observed data sets mainly for the period 1979–2005, we find that anomalous warming events different from conventional El Ni09o events occur in the central equatorial Pacific. This unique warming in the central equatorial Pacific associated with a horseshoe pattern is flanked by a colder sea surface temperature anomaly (SSTA) on both sides along the equator. empirical orthogonal function (EOF) analysis of monthly tropical Pacific SSTA shows that these events are represented by the second mode that explains 12% of the variance. Since a majority of such events are not part of El Ni09o evolution, the phenomenon is named as El Ni09o Modoki (pseudo-El Ni09o) (“Modoki” is a classical Japanese word, which means “a similar but different thing”). The El Ni09o Modoki involves ocean-atmosphere coupled processes which include a unique tripolar sea level pressure pattern during the evolution, analogous to the Southern Oscillation in the case of El Ni09o. Hence the total entity is named as El Ni09o–Southern Oscillation (ENSO) Modoki. The ENSO Modoki events significantly influence the temperature and precipitation over many parts of the globe. Depending on the season, the impacts over regions such as the Far East including Japan, New Zealand, western coast of United States, etc., are opposite to those of the conventional ENSO. The difference maps between the two periods of 1979–2004 and 1958–1978 for various oceanic/atmospheric variables suggest that the recen
    Chen Z. S., Z. P. Wen, R. G. Wu, P. Zhao, and J. Cao, 2014: Influence of two types of El Niños on the East Asian climate during boreal summer: A numerical study. Climate Dyn., 43, 469- 481.30749ba6-bdc8-43d0-abee-5bfd74af7b62141e861fb14f63cf41aa32026f15eac6http%3A%2F%2Flink.springer.com%2F10.1007%2Fs00382-013-1943-1refpaperuri:(e7d4d9010e912477decc8666ee57ed5c)/s?wd=paperuri%3A%28e7d4d9010e912477decc8666ee57ed5c%29&filter=sc_long_sign&tn=SE_xueshusource_2kduw22v&sc_vurl=http%3A%2F%2Flink.springer.com%2F10.1007%2Fs00382-013-1943-1&ie=utf-8&sc_us=6744734652566482819
    Feng J., J. P. Li, 2011: Influence of El Niño Modoki on spring rainfall over south China. J. Geophys. Res.: Atmos., 116, D13102.10.1029/2010JD015160e667fa0f-2a4e-4f46-a974-0ddc6f51376143e3316bb6f806cb7c77b73dc4cf788fhttp%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1029%2F2010JD015160%2Fpdf%2Fenhancedrefpaperuri:(4d9d878ed43b03a2c6471e57da42e5c3)http://onlinelibrary.wiley.com/doi/10.1029/2010JD015160/pdf/enhancedUsing observed data sets from 1979 to 2006, the relationship between El Ni09o Modoki and spring rainfall over south China (SC) is investigated. Of particular interest is the difference in the influence on spring rainfall of typical El Ni09o events and the recently recognized El Ni09o Modoki events, which are characterized by distinct warm sea surface temperature anomalies (SSTA) in the central Pacific and weaker cold anomalies in the western and eastern parts of the basin. Associated with the SSTA, anomalous ascent occurs over the central Pacific and downward flow is observed over the eastern and western Pacific. The anomalous flow is associated with anomalous convergence in the upper troposphere over the western Pacific. SC is influenced by an anomalous anticyclonic circulation with prevailing northeasterly anomalies. The convective activity in SC becomes weaker, resulting in reduced rainfall. However, the situation is different in the case of El Ni09o, in terms of the influence on rainfall over SC. While El Ni09o Modoki events are accompanied by a significant reduction in rainfall over SC, there is enhanced rainfall associated with El Ni09o events. Moreover, there exists a strong asymmetry in the relationship between SC spring rainfall, typical El Ni09o-Southern Oscillation (ENSO) and ENSO Modoki events. It appears that these relationships are only statistically significant for positive events. The asymmetric influence of positive and negative in two ENSO phenomena may explain the difference in their respective relationships with spring rainfall over SC.
    Gushchina D., B. Dewitte, 2012: Intraseasonal tropical atmospheric variability associated with the two flavors of El Niño. Mon. Wea. Rev., 140, 3669- 3681.
    Hendon H. H., C. D. Zhang, and J. D. Glick, 1999: Interannual variation of the Madden-Julian oscillation during austral summer. J.Climate, 12, 2538- 2550.be742f50327233469489774bf66f7378http%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F1999jcli...12.2538h/s?wd=paperuri%3A%28ec235333f148c8c634fa5ff017bd4df5%29&filter=sc_long_sign&tn=SE_xueshusource_2kduw22v&sc_vurl=http%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F1999jcli...12.2538h&ie=utf-8&sc_us=6593899536110148732
    Hendon H. H., M. C. Wheeler, and C. D. Zhang, 2007: Seasonal dependence of the MJO-ENSO relationship. J.Climate, 20, 531- 543.af246c0d-8be1-42dc-8315-6f8d97f3055e1343dad17627bbe2890aae7e4ec67f1dhttp%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F2007JCli...20..531Hrefpaperuri:(d3ada79167f30598a47b92ddc6963f6e)/s?wd=paperuri%3A%28d3ada79167f30598a47b92ddc6963f6e%29&filter=sc_long_sign&tn=SE_xueshusource_2kduw22v&sc_vurl=http%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F2007JCli...20..531H&ie=utf-8&sc_us=17461309860912955640
    Kao H. Y., J. Y. Yu, 2009: Contrasting eastern-Pacific and central-Pacific types of ENSO. J.Climate, 22, 615- 632.10.1175/2008JCLI2309.145c44667cdecca214fcb46319cfa9a89http%3A%2F%2Fwww.cabdirect.org%2Fabstracts%2F20093117308.htmlhttp://www.cabdirect.org/abstracts/20093117308.htmlNot Available
    Kessler W. S., 2001: EOF representations of the Madden-Julian Oscillation and its connection with ENSO. J.Climate, 14, 3055- 3061.10.1175/1520-0442(2001)014<3055:EROTMJ>2.0.CO;2b35f071a-5f22-4a7c-b646-95fa72b12044142a07d5f452ca9f446be39e2c1ca04ahttp%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F2001JCli...14.3055Krefpaperuri:(e48ae6613a7c1a8015d07ead06122668)http://adsabs.harvard.edu/abs/2001JCli...14.3055KAlthough recent El Niño events have seen the occurrence of strong intraseasonal winds apparently associated with the Madden-Julian oscillation (MJO), the usual indices of interannual variability of the MJO are uncorrelated with measures of the ENSO cycle. An EOF decomposition of intraseasonal outgoing longwave radiation and zonal wind identifies two modes of interannual variability of the MJO: a zonally stationary variation of amplitude that is unrelated to ENSO and a roughly 20°-longitude eastward extension of the MJO envelope during El Niño events. The stationary mode is represented by the first two EOFs, which form the familiar lag-correlated quadrature pair, and the eastward-extending mode is represented by the third EOF, which is usually ignored although it is statistically significant. However, the third EOF also has a systematic phase relation with the first pair, and all three should be considered as a triplet; rotating the EOFs makes the phase relation clear. The zonal shift represents about 20% of total MJO variance (which itself is about 55% of intraseasonal variance over the tropical strip). Although the eastward shift is small when compared with the global scale of the MJO, it produces a large proportional shift of MJO activity over the open Pacific, where physical interactions with ENSO processes can occur.
    Kessler W. S., R. Kleeman, 2000: Rectification of the Madden-Julian Oscillation into the ENSO cycle. J.Climate, 13, 3560- 3575.10.1175/1520-0442(2000)013<3560:ROTMJO>2.0.CO;245699f178b60a7b99145753edaaa1f26http%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F2000JCli...13.3560Khttp://adsabs.harvard.edu/abs/2000JCli...13.3560KAn ocean general circulation model, forced with idealized, purely oscillating wind stresses over the western equatorial Pacific similar to those observed during the Madden-Julian oscillation (MJO), developed rectified low-frequency anomalies in SST and zonal currents, compared to a run in which the forcing was climatological. The rectification in SST resulted from increased evaporation under stronger than normal winds of either sign, from correlated intraseasonal oscillations in both vertical temperature gradient and upwelling speed forced by the winds, and from zonal advection due to nonlinearly generated equatorial currents. The net rectified signature produced by the MJO-like wind stresses was SST cooling (about 0.4°C) in the west Pacific, and warming (about 0.1°C) in the central Pacific, tending to flatten the background zonal SST gradient. It is hypothesized that, in a coupled system, such a pattern of SST anomalies would spawn additional westerly wind anomalies as a result of SST-induced changes in the low-level zonal pressure gradient. This was tested in an intermediate coupled model initialized to 1 January 1997, preceding the 1997-98 El Niño. On its own, the model hindcast a relatively weak warm event, but when the effect of the rectified SST pattern was imposed, a coupled response produced the hypothesized additional westerlies and the hindcast El Niño became about 50% stronger (measured by east Pacific SST anomalies), suggesting that the MJO can interact constructively with the ENSO cycle. This implies that developing the capacity to predict, if not individual MJO events, then the conditions that affect their amplitude, may enhance predictability of the strength of oncoming El Niños.
    Kim S. T., J. Y. Yu, 2012: The two types of ENSO in CMIP5 models. Geophys. Res. Lett., 39, L11704.10.1029/2012GL052006b4418360ec3a2a6a6025d2d82813920bhttp%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1029%2F2012GL052006%2Ffullhttp://onlinelibrary.wiley.com/doi/10.1029/2012GL052006/fullIn this study, we evaluate the intensity of the Central-Pacific (CP) and Eastern-Pacific (EP) types of El Niño-Southern Oscillation (ENSO) simulated in the pre-industrial, historical, and the Representative Concentration Pathways (RCP) 4.5 experiments of the Coupled Model Intercomparison Project Phase 5 (CMIP5). Compared to the CMIP3 models, the pre-industrial simulations of the CMIP5 models are found to (1) better simulate the observed spatial patterns of the two types of ENSO and (2) have a significantly smaller inter-model diversity in ENSO intensities. The decrease in the CMIP5 model discrepancies is particularly obvious in the simulation of the EP ENSO intensity, although it is still more difficult for the models to reproduce the observed EP ENSO intensity than the observed CP ENSO intensity. Ensemble means of the CMIP5 models indicate that the intensity of the CP ENSO increases steadily from the pre-industrial to the historical and the RCP4.5 simulations, but the intensity of the EP ENSO increases from the pre-industrial to the historical simulations and then decreases in the RCP4.5 projections. The CP-to-EP ENSO intensity ratio, as a result, is almost the same in the pre-industrial and historical simulations but increases in the RCP4.5 simulation.
    Knutson T. R., K. M. Weickmann, and J. E. Kutzbach, 1986: Global-scale intraseasonal oscillations of outgoing longwave radiation and 250 mb zonal wind during Northern Hemisphere summer. Mon. Wea. Rev., 114, 605- 623.07afcac80d589c1cc58642dd490826eahttp%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F1986MWRv..114..605K/s?wd=paperuri%3A%28df9fe1a390b0380e5b763a1cd2063fae%29&filter=sc_long_sign&tn=SE_xueshusource_2kduw22v&sc_vurl=http%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F1986MWRv..114..605K&ie=utf-8&sc_us=8996603815436954705
    Lafleur D. M., B. S. Barrett, and G. R. Henderson, 2015: Some climatological aspects of the Madden-Julian Oscillation (MJO). J.Climate, 28, 6039- 6053.10.1175/JCLI-D-14-00744.1fe9fc36708498c5418d48cd06c314d7dhttp%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F2015JCli...28.6039Lhttp://adsabs.harvard.edu/abs/2015JCli...28.6039LNot Available
    Larkin N. K., D. E. Harrison, 2005: On the definition of El Niño and associated seasonal average U.S. weather anomalies. Geophys. Res. Lett., 32, L13705.10.1029/2005GL022738e310b5b3-bec0-4fd6-b057-6e543ae87ea53ea54c0835eedec161f7d045840ba6d3http%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1029%2F2005GL022738%2Fcitedbyrefpaperuri:(fc1bc2b52c89e864194ad74f0ff9d71f)http://onlinelibrary.wiley.com/doi/10.1029/2005GL022738/citedbyA new NOAA definition of El Ni09o identifies a number of additional El Ni09o seasons beyond those conventionally agreed. These additional seasons are characterized by SST anomalies primarily in the western central equatorial Pacific. We show here that the seasonal weather anomalies over the U.S. associated with these additional Dateline El Ni09o seasons are substantially different from those associated with conventional El Ni09o seasons. Although some regions have similar associated anomalies, most of the major regional anomalies are quite different. Treating the two as a single phenomenon yields weaker overall seasonal weather associations and does not take advantage of the stronger associations available when the two are treated separately.
    Lau K. M., P. H. Chan, 1986: The 40-50 day oscillation and the El Niño/Southern Oscillation: A new perspective. Bull. Amer. Meteor. Soc., 67, 533- 534.10.1175/1520-0477(1986)067<0533:TDOATE>2.0.CO;2b8afbeeb-3867-4823-afd2-650f564c6656ea7b1d17d4c6a9aa194ea1d16fe27e83http%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F1986BAMS...67..533Lrefpaperuri:(b334477b2ec77ff90d1cecfa6f9ecbe8)http://adsabs.harvard.edu/abs/1986BAMS...67..533LThe tropical Ocean-atmosphere exhibits two prominent modes of low-frequency oscillations, i.e., the "40-50" day oscillation and the El Niño/Southern Oscillation (ENSO). The two phenomena are viewed in the same perspective from 10 years of satellite-derived out-going-longwave-radiation data. Results reveal some interesting features that may lead to new insights into the understanding of the two phenomena.
    Li C. Y., Y. P. Zhou, 1994: Relationship between intraseasonal oscillation in the tropical atmosphere and ENSO. Acta Geophysica Sinica, 37, 17- 26. (in Chinese)
    Madden R. A., P. R. Julian, 1971: Detection of a 40-50 day oscillation in the zonal wind in the tropical Pacific.. J. Atmos. Sci, 28, 702- 708.10.1175/1520-0469(1971)0282.0.CO;23283107661179653056216429222322221870922181624324158132064393089c530a2-c983-404c-ae03-f02037a778eaa33db74c6b65a47f5bd8c4d77c78e938http%3A%2F%2Fwww.cabdirect.org%2Fabstracts%2F20093346088.htmlrefpaperuri:(3a107661c1aa7f9e6ad53ca0562d1b64)http://www.cabdirect.org/abstracts/20093346088.htmlAbstract Nearly ten years of daily rawinsonde data for Canton Island (3S, 172W) have been subjected to spectrum and cross-spectrum analysis. In the course of this analysis a very pronounced maximum was noted in the co-spectrum of the 850- and 150-mb zonal wind components in the frequency range 0.0245–0.0190 day 611 (41–53 days period). Application of a posteriori sampling theory resulted in a significance level of 656% (0.1% prior confidence level). This type of significance test is appropriate because no prior evidence or reason existed for expecting such a spectral feature. Subsequent analysis revealed the following structure of the oscillation. Peaks in the variance spectra of the zonal wind are strong in the low troposphere, are weak or non-existent in the 700–400 mb layer, and are strong again in the upper troposphere. No evidence of this feature could be found above 80 mb, or in any of the spectra of the meridional component. The spectrum of station pressure possesses a peak in this frequency range and the oscillation is in phase with the low tropospheric zonal wind oscillation, and out of phase with that in the upper troposphere. The tropospheric temperatures exhibit a similar peak and are highly coherent with the station pressure oscillation; positive station pressure anomalies are associated with negative temperature anomalies throughout the troposphere. Thus, the lower-middle troposphere appears to be a nodal surface with u and P oscillating in phase but 180° out of phase above and below this surface. Evidence for this phenomenon was found in shorter records at Kwajalein (9N, 168E) but not at Singapore (1N, 104E) or Balboa, Canal Zone (9N, 79w). We speculate that the oscillation is a large circulation cell oriented in zonal planes and centered in the mid-Pacific.
    Madden R. A., P. R. Julian, 1972: Description of global-scale circulation cells in the tropics with a 40-50 day period. J. Atmos. Sci., 29, 1109- 1123.10.1175/1520-0469(1972)029<1109:DOGSCC>2.0.CO;20ad16bc86e058808314db803cb9e6916http%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F1972jats...29.1109mhttp://adsabs.harvard.edu/abs/1972jats...29.1109mCiteSeerX - Scientific documents that cite the following paper: Description of global-scale circulation cells in the tropics with a 40–50 day period
    Madden R. A., P. R. Julian, 1994: Observations of the 40-50-day tropical oscillation- review. Mon. Wea. Rev., 122, 814- 837.
    Marshall A. G., H. H. Hendon, and G. M. Wang, 2016: On the role of anomalous ocean surface temperatures for promoting the record Madden-Julian Oscillation in March 2015. Geophys. Res. Lett., 43, 472- 481.10.1002/2015GL0669841e63db220a600b2fcee88a7c7d0ecafehttp%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F2016GeoRL..43..472Mhttp://adsabs.harvard.edu/abs/2016GeoRL..43..472MA Madden-Julian Oscillation (MJO) event dramatically amplified at the beginning of March 2015 as the convective phase traversed an unusually warm central Pacific Ocean. This record amplification also resulted in record amplitude of the MJO based on index measurements since 1974. We explore the possible role of the anomalously high ocean surface temperatures in the equatorial central Pacific for promoting the extraordinary amplification of this MJO event. Forecast sensitivity experiments with the Predictive Ocean Atmosphere Model for Australia show that the enhanced growth of the MJO resulted from amplification of the convective anomaly as it encountered the unusually warm central Pacific. Our results indicate that anomalous sea surface temperature (SST) at the onset of El Niño 2015 promoted the intensification of the MJO. We suggest a two-way interaction whereby initial SST anomalies promoted enhanced MJO activity which then possibly led to enhanced El Niño development.
    Pohl B., A. J. Matthews, 2007: Observed changes in the lifetime and amplitude of the Madden-Julian Oscillation associated with interannual ENSO sea surface temperature anomalies. J.Climate, 20, 2659- 2674.10.1175/JCLI4230.1bd930db0-0482-4db8-b3b1-21edef54aa7fa359afa17bfc0968e5d12944842b20eehttp%3A%2F%2Fwww.researchgate.net%2Fpublication%2F43033212_Observed_changes_in_the_life_time_and_amplitude_of_the_madden-julian_oscillation_associated_with_interannual_ENSO_sea_surface_temperature_anomaliesrefpaperuri:(d17c66fbd54a172a1bf33f852e1a5695)http://www.researchgate.net/publication/43033212_Observed_changes_in_the_life_time_and_amplitude_of_the_madden-julian_oscillation_associated_with_interannual_ENSO_sea_surface_temperature_anomaliesAbstract The Madden–Julian oscillation (MJO) is analyzed using the reanalysis zonal wind– and satellite outgoing longwave radiation–based indices of Wheeler and Hendon for the 1974–2005 period. The average lifetime of the MJO events varies with season (36 days for events whose central date occurs in December, and 48 days for events in September). The lifetime of the MJO in the equinoctial seasons (March–May and October–December) is also dependent on the state of El Ni09o–Southern Oscillation (ENSO). During October–December it is only 32 days under El Ni09o conditions, increasing to 48 days under La Ni09a conditions, with similar values in northern spring. This difference is due to faster eastward propagation of the MJO convective anomalies through the Maritime Continent and western Pacific during El Ni09o, consistent with theoretical arguments concerning equatorial wave speeds. The analysis is extended back to 1950 by using an alternative definition of the MJO based on just the zonal wind component of the Wheeler and Hendon indices. A rupture in the amplitude of the MJO is found in 1975, which is at the same time as the well-known rupture in the ENSO time series that has been associated with the Pacific decadal oscillation. The mean amplitude of the MJO is 16% larger in the postrupture (1976–2005) compared to the prerupture (1950–75) period. Before the 1975 rupture, the amplitude of the MJO is maximum (minimum) under El Ni09o (La Ni09a) conditions during northern winter, and minimum (maximum) under El Ni09o (La Ni09a) conditions during northern summer. After the rupture, this relationship disappears. When the MJO–ENSO relationship is analyzed using all-year-round data, or a shorter dataset (as in some previous studies), no relationship is found.
    Straub K. H., 2013: MJO initiation in the real-time multivariate MJO index. J.Climate, 26, 1130- 1151.10.1175/JCLI-D-12-00074.1651d5a39cf147a3a50a139c47066e2dchttp%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F2013JCli...26.1130Shttp://adsabs.harvard.edu/abs/2013JCli...26.1130SAbstract Madden–Julian oscillation (MJO) initiation in the real-time multivariate MJO (RMM) index is explored through an analysis of observed case studies and composite events. Specific examples illustrate that both the dates of MJO initiation and the existence of the MJO itself can vary substantially among several well-known MJO indices, depending on whether the focus is on convection or circulation. Composites of “primary” MJO initiation events in which the RMM index rapidly increases in amplitude from a non-MJO state to an MJO state are presented and are supplemented by two case studies from the 1985/86 winter season. Results illustrate that, for primary MJO initiation events in the Indian Ocean (RMM phase 1), slowly eastward-propagating 850-hPa (200 hPa) easterly (westerly) anomalies over the Indian Ocean precede the amplification of the RMM index by at least 10 days, while suppressed convection over the western Pacific Ocean precedes the amplification by 5 days. These “local” Eastern Hemispheric predecessor signals are similar to those found in successive (well established) MJO events but are not captured by the global-scale RMM index because of their smaller zonal scale. The development of a primary MJO event is thus often transparent in the RMM index, since it occurs on scales smaller than zonal wavenumber 1, particularly in convection. Even when the RMM index is altered to respond to convection only, the same local precursor signals are found. Both composites and case studies suggest that, for primary MJO initiation events in the Indian Ocean, the development of global-scale circulation anomalies typically precedes the onset of large-scale deep convection.
    Tam C. Y., N. C. Lau, 2005: Modulation of the Madden-Julian Oscillation by ENSO: Inferences from observations and GCM simulations. J. Meteor. Soc.Japan, 83, 727- 743.10.2151/jmsj.83.727941b6ab737c7931068431722e1e93d92http%3A%2F%2Fci.nii.ac.jp%2Fnaid%2F130004788511http://ci.nii.ac.jp/naid/130004788511The impact of the El Nino-Southern Oscillation (ENSO) on the Madden-Julian Oscillation (MJO) is studied, based on reanalysis data and output from an ensemble general circulation model (GCM) experiment. Observed monthly sea surface temperature variations over the period of 1950-99 are imposed in the deep tropical eastern/central Pacific in the course of the SST experiment. Both GCM, and reanalysis data, indicate that intraseasonal activity of the low-level zonal wind is enhanced (reduced) over the central (western) Pacific during El Nino events. The propagation and growth/decay characterisitcis of the MJO in different phases of ENSO is also examined, based on a lag correlation technique. During warm events there is an eastward shift in the locations of strong growth and decay, and the propagation of the MJO becomes slower in the warm ENSO phase. These changes are reversed during La Nina epsiodes. Using output from the GCM experiment, the effects of ENSO on the circulation and convection during the MJO lifecycle are studied in detail. Further eastward penetration of MJO-related convection is simulated during warm events over the central Pacific. An instability index related to the vertical gradient of the moist static energy is found to be useful for depicting the onset of MJO convection along the equator. During warm events, the stronger magnitudes of this index over the central Pacific are conducive to more eastward penetration of convective anomalies in the region. These changes are mainly due to the intensified moisture accumulation at low levels. Analysis of the moisture budget suggests that the stronger moisture accumulation can be related to the increased low-level humidity over the central Pacific during warm events.
    Trenberth K. E., D. P. Stepaniak, 2001: Indices of El Niño evolution. J.Climate, 14, 1697- 1701.10.1175/1520-0442(2001)014<1697:LIOENO>2.0.CO;2c49ddf9b-e9ce-41e3-86a0-38934a7332d7ef097288f9b217dc7db4cdb08c5cd402http%3A%2F%2Fwww.researchgate.net%2Fpublication%2F256476464_Indices_of_El_Nio_Evolutionrefpaperuri:(ff5364e3e0aaf688a2397f152836678b)http://www.researchgate.net/publication/256476464_Indices_of_El_Nio_EvolutionAbstract To characterize the nature of El Ni09o–Southern Oscillation (ENSO), sea surface temperature (SST) anomalies in different regions of the Pacific have been used. An optimal characterization of both the distinct character and the evolution of each El Ni09o or La Ni09a event is suggested that requires at least two indices: (i) SST anomalies in the Ni09o-3.4 region (referred to as N3.4), and (ii) a new index termed here the Trans-Ni09o Index (TNI), which is given by the difference in normalized anomalies of SST between Ni09o-1+2 and Ni09o-4 regions. The first index can be thought of as the mean SST throughout the equatorial Pacific east of the date line and the second index is the gradient in SST across the same region. Consequently, they are approximately orthogonal. TNI leads N3.4 by 3 to 12 months prior to the climate shift in 1976/77 and also follows N3.4 but with opposite sign 3 to 12 months later. However, after 1976/77, the sign of the TNI leads and lags are reversed.
    Trenberth K. E., D. P. Stepaniak, and J. M. Caron, 2002: Interannual variations in the atmospheric heat budget. J. Geophys. Res.,107(D8), AAC4-1-ACC4-15.10.1029/2000JD00029785e761b36c2ec03e2142cc6dc5a02054http%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1029%2F2000JD000297%2Fcitedbyhttp://onlinelibrary.wiley.com/doi/10.1029/2000JD000297/citedbyInterannual variability of the atmospheric heat budget is explored via a new data set of the computed vertically integrated energy transports to examine relationships with other fields. A case study reveals very large monthly divergences of these transports regionally with El Niño-Southern Oscillation (ENSO) and the associated changes with the Pacific-North American teleconnection pattern, and with the North Atlantic Oscillation. In the tropical Pacific during large El Niño events the anomalous divergence of the atmospheric energy transports exceeds 50 W mover broad regions for several months. Examination of the corresponding top-of-the-atmosphere net radiative fluxes shows that it is primarily the surface fluxes from the ocean to the atmosphere that feed the divergent atmospheric transports. A systematic investigation of the covariability of sea surface temperatures (SSTs) and the divergence of atmospheric energy transport, using singular value decomposition analysis of the temporal covariance, reveals ENSO as dominant in the first two modes, explaining 62% and 12% of the covariance in the Pacific domain and explaining 39.5% and 15.4% globally for the first and second modes, respectively. The first mode is well represented by the time series for the SST index for Niño 3.4 region (170°W-120°W, 5°N-5°S). Regression analysis allows a more complete view of how the SSTs, outgoing longwave radiation, precipitation, diabatic heating, and atmospheric circulation respond with ENSO. The second mode indicates aspects of the systematic evolution of ENSO with time, with strong lead and lag correlations. It primarily reflects differences in the evolution of ENSO across the tropical Pacific from about the dateline to coastal South America. High SSTs associated with warm ENSO events are damped through surface heat fluxes into the atmosphere, which transports the energy into higher latitudes and throughout the tropics, contributing to loss of heat by the ocean, while the cold ENSO events correspond to a recharge phase as heat enters the ocean. Diabatic processes are clearly important within ENSO evolution.
    Weickmann K. M., G. R. Lussky, and J. E. Kutzbach, 1985: Intraseasonal (30-60 day) fluctuations of outgoing longwave radiation and 250 mb streamfunction during northern winter. Mon. Wea. Rev., 113, 941- 961.10.1175/1520-0493(1985)113<0941:IDFOOL>2.0.CO;20649d2c5-6f3e-4dd5-82c6-007364525c4d1de87b79c5f97615ae8beb72e6cc06b6http%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F1985MWRv..113..941Wrefpaperuri:(5bc420765b507637ce139f81ca89d1e5)http://adsabs.harvard.edu/abs/1985MWRv..113..941WNot Available
    Weng H. Y., K. Ashok, S. K. Behera, S. A. Rao, and T. Yamagata, 2007: Impacts of recent El Niño Modoki on dry/wet conditions in the Pacific Rim during boreal summer. Climate Dyn., 29, 113- 129.10.1007/s00382-007-0234-0ce13b59d-219f-46ee-8bd6-c00d07dd7bd5d8d1f09734699318f8a5bb034c67b416http%3A%2F%2Fwww.springerlink.com%2Fcontent%2Fa36w326j63468186refpaperuri:(e6a424fe52f801dc11c92e5ab8c3c6ff)http://www.springerlink.com/content/a36w326j63468186Present work uses 1979–2005 monthly observational data to study the impacts of El Ni09o Modoki on dry/wet conditions in the Pacific rim during boreal summer. The El Ni09o Modoki phenomenon is characterized by the anomalously warm central equatorial Pacific flanked by anomalously cool regions in both west and east. Such zonal SST gradients result in anomalous two-cell Walker Circulation over the tropical Pacific, with a wet region in the central Pacific. There are two mid-tropospheric wave trains passing over the extratropical and subtropical North Pacific. They contain a positive phase of a Pacific-Japan pattern in the northwestern Pacific, and a positive phase of a summertime Pacific-North American pattern in the northeastern Pacific/North America region. The western North Pacific summer monsoon is enhanced, while the East Asian summer monsoon is weakened. In the South Pacific, there is a basin-wide low in the mid-latitude with enhanced Australian high and the eastern South Pacific subtropical high. Such an atmospheric circulation pattern favors a dry rim surrounding the wet central tropical Pacific. The El Ni09o Modoki and its climate impacts are very different from those of El Ni09o. Possible geographical regions for dry/wet conditions influenced by El Ni09o Modoki and El Ni09o are compared. The two phenomena also have very different temporal features. El Ni09o Modoki has a large decadal background while El Ni09o is predominated by interannual variability. Mixing-up the two different phenomena may increase the difficulty in understanding their mechanisms, climate impacts, and uncertainty in their predictions.
    Wheeler M. C., H. H. Hendon, 2004: An all-season real-time multivariate MJO index: Development of an index for monitoring and prediction. Mon. Wea. Rev., 132, 1917- 1932.ff3949520e2db3649691459062f6df4dhttp%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F2004MWRv..132.1917W/s?wd=paperuri%3A%280c747c09ff298b439f3e819ddb6c0cdb%29&filter=sc_long_sign&tn=SE_xueshusource_2kduw22v&sc_vurl=http%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F2004MWRv..132.1917W&ie=utf-8&sc_us=17249076170115512361
    Yasunari T., 1980: A quasi-stationary appearance of 30 to 40 day period in the cloudiness fluctuations during the summer monsoon over India. J. Meteor. Soc.Japan, 58, 225- 229.825a9ed4-7b66-4971-bba8-fb164cb5310e40887e4f22e3ec3aeb9cf695d834b576http%3A%2F%2Fciteseer.uark.edu%3A8080%2Fciteseerx%2Fshowciting%3Bjsessionid%3D8EEB9A7135806F5FF0A6381AB31E33E9%3Fcid%3D9212567refpaperuri:(0013f28477ffdbfc8bf9ac6533e70190)http://citeseer.uark.edu:8080/citeseerx/showciting;jsessionid=8EEB9A7135806F5FF0A6381AB31E33E9?cid=9212567CiteSeerX - Scientific documents that cite the following paper: A quasi-stationary appearance of 30-40 day period in the cloudiness fluctuations during the summer monsoon over India
    Yeh S. W., J. S. Kug, B. Dewitte, M. H. Kwon, B. P. Kirtman, and F. F. Jin, 2009: El Niño in a changing climate. Nature, 461, 511- 514.10.1038/nature08316cbeb8c90-b572-4447-b285-437cfa49917348ff6213f3ddfb718f6b8dbdaa37ea1chttp%3A%2F%2Feuropepmc.org%2Fabstract%2FMED%2F19779449refpaperuri:(8b689bbb288d5528989f50569ccb2536)http://europepmc.org/abstract/MED/19779449El Ni09o events, characterized by anomalous warming in the eastern equatorial Pacific Ocean, have global climatic teleconnections and are the most dominant feature of cyclic climate variability on subdecadal timescales. Understanding changes in the frequency or characteristics of El Ni09o events in a changing climate is therefore of broad scientific and socioeconomic interest. Recent studies show that the canonical El Ni09o has become less frequent and that a different kind of El Ni09o has become more common during the late twentieth century, in which warm sea surface temperatures () in the central Pacific are flanked on the east and west by cooler . This type of El Ni09o, termed the central Pacific El Ni09o (CP-El Ni09o; also termed the dateline El Ni09o, El Ni09o Modoki or warm pool El Ni09o), differs from the canonical eastern Pacific El Ni09o (EP-El Ni09o) in both the location of maximum anomalies and tropical-midlatitude teleconnections. Here we show changes in the ratio of CP-El Ni09o to EP-El Ni09o under projected global warming scenarios from the Coupled Model Intercomparison Project phase 3 multi-model data set. Using calculations based on historical El Ni09o indices, we find that projections of anthropogenic climate change are associated with an increased frequency of the CP-El Ni09o compared to the EP-El Ni09o. When restricted to the six climate models with the best representation of the twentieth-century ratio of CP-El Ni09o to EP-El Ni09o, the occurrence ratio of CP-El Ni09o/EP-El Ni09o is projected to increase as much as five times under global warming. The change is related to a flattening of the thermocline in the equatorial Pacific.
    Yuan Y., C. Y. Li, and J. Ling, 2015: Different MJO activities between EP El Niño and CP El Niño. Scientia Sinica Terrae, 45, 318- 334. (in Chinese)
    Yuan Y., H. Yang, and C. Y. Li, 2014: Possible influences of the tropical Indian Ocean dipole on the eastward propagation of MJO. Journal of Tropical Meteorology, 20, 173- 180.10.1016/j.asr.2014.02.02381bb2d29dc5cf8b0cce4205b2f241298http%3A%2F%2Fwww.cnki.com.cn%2FArticle%2FCJFDTotal-RQXB201402009.htmhttp://www.cnki.com.cn/Article/CJFDTotal-RQXB201402009.htm
    Zhang C. D., 2005: Madden-Julian oscillation. Rev. Geophys.,43, RG2003.10.1029/2004RG000158579261ee1156a68aeb0437dcd3876dafhttp%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1002%2F0471743984.vse9530%2Fabstract%3Bjsessionid%3DF7AC345CBF75EDA99EA23C2158EDCF0E.f03t02/s?wd=paperuri%3A%28237065d1f4c9d275a6c0e2d43f50b759%29&filter=sc_long_sign&tn=SE_xueshusource_2kduw22v&sc_vurl=http%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1002%2F0471743984.vse9530%2Fabstract%3Bjsessionid%3DF7AC345CBF75EDA99EA23C2158EDCF0E.f03t02&ie=utf-8&sc_us=18130879363342324841ABSTRACT 1] The Madden-Julian Oscillation (MJO) is the dominant component of the intraseasonal (30 - 90 days) variability in the tropical atmosphere. It consists of large-scale coupled patterns in atmospheric circulation and deep convection, with coherent signals in many other variables, all propagating eastward slowly ($5 m s 脌1) through the portion of the Indian and Pacific oceans where the sea surface is warm. It constantly interacts with the underlying ocean and influences many weather and climate systems. The past decade has witnessed an expeditious progress in the study of the MJO: Its large-scale and multiscale structures are better described, its scale interaction is recognized, its broad influences on tropical and extratropical weather and climate are increasingly appreciated, and its mechanisms for disturbing the ocean are further comprehended. Yet we are facing great difficulties in accurately simulating and predicting the MJO using sophisticated global weather forecast and climate models, and we are unable to explain such difficulties based on existing theories of the MJO. It is fair to say that the MJO remains an unmet challenge to our understanding of the tropical atmosphere and to our ability to simulate and predict its variability. This review, motivated by both the acceleration and gaps in our knowledge of the MJO, intends to synthesize what we currently know and what we do not know on selected topics: its observed basic characteristics, mechanisms, numerical modeling, air-sea interaction, and influences on the El Niño and Southern Oscillation.
    Zhang W. J., F. F. Jin, J. P. Li, and H. L. Ren, 2011: Contrasting impacts of two-type El Niño over the western North Pacific during boreal autumn. J. Meteor. Soc.Japan, 89, 563- 569.10.2151/jmsj.2011-510128a4e2a-9725-46cf-a6e5-529fd8d0db57eaf551f8bde5c46c0411b7d4f1efc37fhttp%3A%2F%2Fci.nii.ac.jp%2Fnaid%2F40019052112refpaperuri:(5261285a6ff448f169e4905c3f7937b8)http://ci.nii.ac.jp/naid/40019052112This work contrasts the climatic impacts of so-called warm-pool (WP) and cold-tongue (CT) El Ni09o on the atmospheric circulation over the western North Pacific (WNP). It is found that the anomalous atmospheric circulation over the WNP is nearly opposite in response to these two types of El Ni09o events in developing autumn. A weak anomalous anticyclone appears over the WNP during CT El Ni09o, whereas a weak anomalous cyclone emerges in the same region during WP El Ni09o. These nearly opposite autumn responses of atmospheric circulation have a significant impact on East Asian climate, and southern China autumn rainfall in particular, although this contrast tends to diminish as El Ni09o events enter their mature phase.
  • [1] Yuanpu LI, Wenshou TIAN, 2017: Different Impact of Central Pacific and Eastern Pacific El Niño on the Duration of Sudden Stratospheric Warming, ADVANCES IN ATMOSPHERIC SCIENCES, 34, 771-782.  doi: 10.1007/s00376-017-6286-0
    [2] HUANG Ping, WANG Pengfei, HU Kaiming, HUANG Gang, ZHANG Zhihua, LIU Yong, YAN Bangliang, 2014: An Introduction to the Integrated Climate Model of the Center for Monsoon System Research and Its Simulated Influence of El Nio on East Asian-Western North Pacific Climate, ADVANCES IN ATMOSPHERIC SCIENCES, 31, 1136-1146.  doi: 10.1007/s00376-014-3233-1
    [3] XIE Fei, LI Jianping, TIAN Wenshou, ZHANG Jiankai, SHU Jianchuan, 2014: The Impacts of Two Types of El Nio on Global Ozone Variations in the Last Three Decades, ADVANCES IN ATMOSPHERIC SCIENCES, 31, 1113-1126.  doi: 10.1007/s00376-013-3166-0
    [4] Wansuo DUAN, Chaoming HUANG, Hui XU, 2017: Nonlinearity Modulating Intensities and Spatial Structures of Central Pacific and Eastern Pacific El Niño Events, ADVANCES IN ATMOSPHERIC SCIENCES, 34, 737-756.  doi: 10.1007/s00376-017-6148-9
    [5] Soon-Il AN, 2018: Impact of Pacific Decadal Oscillation on Frequency Asymmetry of El Niño and La Niña Events, ADVANCES IN ATMOSPHERIC SCIENCES, 35, 493-494.  doi: 10.1007/s00376-018-8024-7
    [6] Feng XUE, Xiao DONG, Fangxing FAN, 2018: Anomalous Western Pacific Subtropical High during El Niño Developing Summer in Comparison with Decaying Summer, ADVANCES IN ATMOSPHERIC SCIENCES, 35, 360-367.  doi: 10.1007/s00376-017-7046-x
    [7] N. FREYCHET, S. SPARROW, S.F. B. TETT, M.J. MINETER, G.C. HEGERL, D.C. H. WALLOM, 2018: Impacts of Anthropogenic Forcings and El Niño on Chinese Extreme Temperatures, ADVANCES IN ATMOSPHERIC SCIENCES, 35, 994-1002.  doi: 10.1007/s00376-018-7258-8
    [8] Ben TIAN, Wansuo DUAN, 2016: Comparison of Constant and Time-variant Optimal Forcing Approaches in El Niño Simulations by Using the Zebiak-Cane Model, ADVANCES IN ATMOSPHERIC SCIENCES, 33, 685-694.  doi: 10.1007/s00376-015-5174-8
    [9] LI Jianying, LIU Boqi, LI Jiandong, MAO Jiangyu, 2015: A Comparative Study on the Dominant Factors Responsible for the Weaker-than-expected El Niño Event in 2014, ADVANCES IN ATMOSPHERIC SCIENCES, 32, 1381-1390.  doi: 10.1007/s00376-015-4269-6
    [10] Bin WANG, Juan LI, Qiong HE, 2017: Variable and Robust East Asian Monsoon Rainfall Response to El Niño over the Past 60 Years (1957-2016), ADVANCES IN ATMOSPHERIC SCIENCES, 34, 1235-1248.  doi: 10.1007/s00376-017-7016-3
    [11] Chaofan LI, Wei CHEN, Xiaowei HONG, Riyu LU, 2017: Why Was the Strengthening of Rainfall in Summer over the Yangtze River Valley in 2016 Less Pronounced than that in 1998 under Similar Preceding El Niño Events?——Role of Midlatitude Circulation in August, ADVANCES IN ATMOSPHERIC SCIENCES, 34, 1290-1300.  doi: 10.1007/s00376-017-7003-8
    [12] Chengyang GUAN, Xin WANG, Haijun YANG, 2023: Understanding the Development of the 2018/19 Central Pacific El Niño, ADVANCES IN ATMOSPHERIC SCIENCES, 40, 177-185.  doi: 10.1007/s00376-022-1410-1
    [13] Ngar-Cheung LAUInstitute of Environment, Energy and Sustainability, and Department of Geography and Resource Management, The Chinese University of Hong Kong, Shatin, N.T., Hong Kong, 2017: The Pioneering Works of Professor Duzheng YE on Atmospheric Dispersion, Tibetan Plateau Meteorology, and Air-Sea Interaction, ADVANCES IN ATMOSPHERIC SCIENCES, 34, 1137-1149.  doi: 10.1007/s00376-017-6256-6
    [14] Lin CHEN, Tim LI, Swadhin K. BEHERA, Takeshi DOI, 2016: Distinctive Precursory Air-Sea Signals between Regular and Super El Niños, ADVANCES IN ATMOSPHERIC SCIENCES, 33, 996-1004.  doi: 10.1007/s00376-016-5250-8
    [15] Jian RAO, Rongcai REN, 2017: Parallel Comparison of the 1982/83, 1997/98 and 2015/16 Super El Niños and Their Effects on the Extratropical Stratosphere, ADVANCES IN ATMOSPHERIC SCIENCES, 34, 1121-1133.  doi: 10.1007/s00376-017-6260-x
    [16] Feng XUE, Fangxing FAN, 2016: Anomalous Western Pacific Subtropical High during Late Summer in Weak La Niña Years: Contrast between 1981 and 2013, ADVANCES IN ATMOSPHERIC SCIENCES, 33, 1351-1360.  doi: 10.1007/s00376-016-5281-1
    [17] Xuben LEI, Wenjun ZHANG, Pang-Chi HSU, Chao LIU, 2021: Distinctive MJO Activity during the Boreal Winter of the 2015/16 Super El Niño in Comparison with Other Super El Niño Events, ADVANCES IN ATMOSPHERIC SCIENCES, 38, 555-568.  doi: 10.1007/s00376-020-0261-x
    [18] Zhang Renhe, Zhao Gang, 2001: Meridional Wind Stress Anomalies over the Tropical Pacific and the Onset of El Ni?o Part Ⅱ: Dynamical Analysis, ADVANCES IN ATMOSPHERIC SCIENCES, 18, 1053-1065.  doi: 10.1007/s00376-001-0022-4
    [19] Xiaomeng SONG, Renhe ZHANG, Xinyao RONG, 2019: Influence of Intraseasonal Oscillation on the Asymmetric Decays of El Niño and La Niña, ADVANCES IN ATMOSPHERIC SCIENCES, , 779-792.  doi: 10.1007/s00376-019-9029-6
    [20] Congxi FANG, Yu LIU, Qiufang CAI, Huiming SONG, 2021: Why Does Extreme Rainfall Occur in Central China during the Summer of 2020 after a Weak El Niño?, ADVANCES IN ATMOSPHERIC SCIENCES, 38, 2067-2081.  doi: 10.1007/s00376-021-1009-y

Get Citation+

Export:  

Share Article

Manuscript History

Manuscript received: 17 December 2015
Manuscript revised: 23 March 2016
Manuscript accepted: 12 April 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

Impacts of Two Types of El Niño on the MJO during Boreal Winter

  • 1. Center for Monsoon and Environment Research/School of Atmospheric Sciences, Sun Yat-Sen University, Guangzhou 510275
  • 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. State Key Laboratory of Tropical Oceanography, South China Sea Institute of Oceanology, Chinese Academy of Sciences, Guangzhou 510301

Abstract: The features of the MJO during two types of El Niño events are investigated in this paper using the daily NCEP-2 reanalysis data, OLR data from NOAA, and Real-time Multivariate MJO index for the period 1979-2012. The results indicate that the MJO exhibits distinct features during eastern Pacific (EP) El Niño events, as compared to central Pacific (CP) El Niño events. First, the intensity of the MJO is weakened during EP El Niño winters from the tropical eastern Indian Ocean to the western Pacific, but enhanced during CP El Niño winters. Second, the range of the MJO eastward propagation is different during the two types of El Niño events. During EP El Niño winters, the MJO propagates eastwards to 120°W, but only to 180° during CP El Niño winters. Finally, the frequency in eight phases of the MJO may be affected by the two types of El Niño. Phases 2 and 3 display a stronger MJO frequency during EP El Niño winters, but phases 4 and 5 during CP El Niño winters.

1. Introduction
  • The MJO is a large-scale eastward-propagating circulation in the atmosphere. (Madden and Julian, 1971) found a 40-50-day oscillation when analyzing the zonal wind anomalies of Canton Island. They pointed out that the MJO is characterized by the planetary scale of wavenumber 1 and eastward propagation (Madden and Julian, 1972). (Yasunari, 1980) confirmed that zonal wind has low-frequency oscillation at the time scale of 40-50 days. (Weickmann et al., 1985) and (Knutson et al., 1986) showed that the MJO also has a vertical baroclinic structure. Many subsequent studies have demonstrated that the active area of the MJO is in the South Asia monsoon region, tropical western Pacific and eastern Pacific (Madden and Julian, 1994; Zhang, 2005).

    Since the MJO is the most dominant signal in tropical intraseasonal variability, and ENSO is the major source of interannual variability in the tropics, many previous studies have investigated the interaction between the MJO and ENSO. For example, the MJO can trigger El Niño through sea-air interaction (Lau and Chan, 1986). Before El Niño occurs, MJO activity is greatly enhanced over the tropical western Pacific (Li and Zhou, 1994), suggesting an impact of the MJO on the occurrence of El Niño events. MJO activity in late boreal spring is favorable to the development of El Niño in the subsequent autumn and winter (Hendon et al., 2007). On the other hand, ENSO can also influence the activity of the MJO. During El Niño, MJO amplitude is relatively weak, implying a weakening effect of El Niño on MJO intensity (Li and Zhou, 1994). During warm ENSO episodes, MJO convective activity shifts eastward to the central and eastern Pacific, with decreased intensity across the eastern Indian Ocean and Maritime Continent (Hendon et al., 1999). Moreover, the strong warm SST anomaly in the central Pacific promotes rapid growth of the MJO in the western Pacific convective phases (Marshall et al., 2016). These previous studies indicate a two-way interaction between enhanced MJO activity and the development of El Niño (Kessler and Kleeman, 2000; Marshall et al., 2016). Besides, the MJO extends further east to the central Pacific during El Niño events, which is associated with the warmer SST beneath (Kessler, 2001; Tam and Lau, 2005). Further study suggests that the lifetime of the MJO is also dependent on the state of ENSO: the MJO propagates faster through the Maritime Continent and western Pacific during El Niño (Pohl and Matthews, 2007).

    Recently, a different type of El Niño, characterized by a warm SST anomaly in the central Pacific, has been widely discussed (Larkin and Harrison, 2005; Ashok et al., 2007). There are many terms to describe this phenomenon, such as "dateline El Niño" (Larkin and Harrison, 2005), "El Niño Modoki" (Ashok et al., 2007), and "central Pacific El Niño" (Kao and Yu, 2009). In this study, we name the two types of El Niño events as "eastern Pacific El Niño" (EP El Niño) and "central Pacific El Niño" (CP El Niño), respectively. During EP El Niño, SST, precipitation and wind anomalies all display dipolar patterns; whereas, they all display tripole patterns during CP El Niño (Ashok et al., 2007; Weng et al., 2007; Kao and Yu, 2009). It is important to note that some studies have raised doubt about the independence of CP El Niño events (Trenberth and Stepaniak, 2001; Trenberth et al., 2002); however, the two types of El Niño do seem to exert different influences on both regional climate and global climate via teleconnection (Larkin and Harrison, 2005; Feng and Li, 2011; Chen et al., 2014). Moreover, recent studies have shown that CP El Niño events have occurred more frequently since the beginning of the 1990s, as compared to EP El Niño events (Yeh et al., 2009; Zhang et al., 2011). Some studies even suggest that the frequency of CP El Niño occurrence will keep on increasing in the 21st century (Kim and Yu, 2012). Therefore, it is necessary to examine the different responses of the MJO to the two types of El Niño events.

    However, few studies have thus far compared the MJO's activity between the two types of El Niño events, although many have examined the interaction between the MJO and ENSO. (Hendon et al., 1999) showed that enhanced MJO activity occurs along with an SST anomaly pattern like CP El Niño. (Gushchina and Dewitte, 2012) demonstrated that the MJO is intensified prior to the peak of EP El Niño, while it is increased during the mature and decaying phases of CP El Niño. (Yuan et al., 2015) further showed the seasonal changes of MJO kinetic energy during the two types of El Niño events. However, these studies mainly focused on comparing MJO intensity during the evolution of the two types of El Niño. It remains unclear whether the two types of El Niño have different influences on MJO simultaneously——not only on MJO intensity, but also on its eastward propagation. Besides, previous studies have generally used the zonal wind at 850 hPa to describe the MJO, which may be greatly constrained by the underlying surface.

    The purpose of the present work is to explore whether the MJO's activity——including its intensity, eastward propagation and active phases——is distinct against the background of the two different types of El Niño events. Since the most active MJO events occur in December-February (Wheeler and Hendon, 2004), we focus on studying the differences during this season, i.e., boreal winter. Instead of 850-hPa zonal wind, we use OLR and 200-hPa velocity potential to depict the MJO. The remainder of the paper is organized as follows: Section 2 described the datasets and methods used in this work. Section 3 presents the impacts on MJO intensity during the two types of El Niño winters. Section 4 compares the MJO's eastward propagation, and conclusions are given in section 5.

2. Data and methods
  • The datasets used in this study include:

    (1) Daily NCEP-2 reanalysis horizontal wind data (resolution: 2.5°× 2.5°) from 1979 to 2012;

    (2) Daily OLR (horizontal resolution: 2.5°× 2.5°), provided by NOAA, from 1979 to 2012;

    (3) Monthly SST data (resolution: 1°× 1°), obtained from NOAA, from 1979 to 2012;

    (4) Real-time Multivariate MJO (RMM) index (Wheeler and Hendon, 2004), obtained from the Australia Meteorological Bureau (http://www.bom.gov.au/climate/mjo/graphics/rmm.74toRealtime.txt), from 1979 to 2012.

    Boreal winter in this paper is defined as the period from December to February. For the sake of simplicity, we use the year of December to represent the year for a particular winter. For example, the "1979 winter" indicates the period from December 1979 to February 1980. All the daily data are dealt to 365 days in each year, which means the data on 29th February in leap years are removed.

    The 30-60-day filtered OLR and 200-hPa velocity potential data are utilized to depict the spatial pattern of the MJO. In particular, the intensity of the MJO is quantified by the variance of these two variables. The RMM index is used to describe the phases of propagation. This index is based on a multivariable EOF analysis of daily OLR, 200-hPa and 850-hPa zonal wind anomalies. The principal components of the first two EOFs (RMM1 and RMM2) can be plotted on a phase-space diagram. It is generally divided into eight phases, and each phase corresponds to a particular stage of the MJO life cycle.

  • To separate the characteristics of the MJO during the two types of El Niño events, the Butterworth bandpass filter and composite analysis are used. An F-test is used to compute the confidence level for the composite of variance anomaly. The degrees of freedom are n-2, where n is the number of cases.

    Following former studies (e.g., Ashok et al., 2007; Weng et al., 2007), the Niño3 index and El Niño Modoki index (EMI) are used to classify EP El Niño and CP El Niño events: Niño3 index is defined as the mean SST anomaly averaged over the equatorial eastern Pacific [(5°S-5°N, 150°-90°W)]; and \begin{equation} {EMI}={SSTA}_{C}-0.5{SSTA}_{E}-0.5{SSTA}_{W} , (1)\end{equation} where SSTA C, SSTA E, and SSTA W represent the area-mean SST anomaly, averaged over the central Pacific [(10°S-10°N, 165°E-140°W)], eastern Pacific [(15°S-5°N, 110°-70°W)] and western Pacific [(10°S-20°N, 125°-145°E)], respectively.

    Figure 1 shows the standardized Niño3 index and EMI in boreal winter from 1979 to 2011. A typical EP (CP) El Niño event is defined when the Niño3 index (EMI) is greater than or equal to one standard deviation, which is represented by the dotted line in Fig. 1. There are two years (1991 and 2009) that meet the criterion of both indices, and thus they are not taken into consideration in this study. Based on the above criteria, there are three EP El Niño years (1982, 1986 and 1997) and four CP El Niño years (1990, 1994, 2002 and 2004).

    Figure 1.  Standardized Niño3 index (black line) and EMI (red line) averaged during boreal winter.

    Figure 2.  Composite anomalies of 30-60-day OLR variance (units: W$^2$ m$^-4$; color-shaded) and SST (units: $^\circ$C; red contours) during (a) EP El Niño winters and (b) CP El Niño winters, and (c) the differences in the 30-60-day OLR variance between EP and CP El Niño winters (dots indicate regions that are statistically significant at the 99% confidence level).

3. Comparison of MJO intensity during the two types of El Niño
  • In order to quantify the MJO intensity, three categories of index——cloudiness, dynamical, and combined cloudiness and dynamical——were generalized by (Straub, 2013). In this study, we utilize OLR data as the cloudiness index and upper-tropospheric zonal winds as dynamical indices to explore the differences of the MJO in response to the two types of El Niño events. The variance of 30-60-day OLR and 200-hPa velocity potential in the tropics are calculated to identify the MJO intensity.

    Figure 2 shows the distribution of the variance anomaly of 30-60-day OLR, and the SST anomaly, which is represented by the contour lines of 1°C and 2°C, during the two types of El Niño winters. The differences between EP and CP El Niño winters are also presented. In the EP El Niño winters (Fig. 2a), the negative variance anomaly appears over the west of the dateline, including the Indian Ocean and the western Pacific, and the positive one appears over the east of the dateline. The negative center lies over the Maritime Continent, while the positive center lies over the central Pacific (near 130°W), which agrees with the results of (Hendon et al., 1999). By contrast, for the CP El Niño winters (Fig. 2b), the positive anomalies appear over the west of the Maritime Continent and the central Pacific near the dateline. Moreover, the warm SST anomaly corresponds to the enhanced MJO convective anomaly during both types of El Niño winters. The difference between EP and CP El Niño events (Fig. 2c) is significantly negative from the tropical eastern Indian Ocean to the western Pacific, which exceeds the 99% confidence level. Thus, it can be concluded that the intensity of 30-60-day OLR from the tropical eastern Indian Ocean to the western Pacific is weaker during EP El Niño winters, while it is stronger during CP winters.

    Figure 3.  As in Fig. 2 but for velocity potential (units: 10$^-12$ m$^4$ s$^-2$) at 200 hPa.

    Figure 3 shows the variance anomaly of 30-60-day velocity potential at 200 hPa during the two types of El Niño winters. The negative variance anomaly occurs over almost the whole of the tropics, and the strongest negative center lies over the eastern Indian Ocean during EP El Niño winters (Fig. 3a). However, when CP El Niño occurs (Fig. 3b), three positive centers of variance anomaly appear over the tropical ocean. The strongest two lie over the eastern Indian Ocean and western Pacific, respectively. A weak negative anomaly over the tropical north-central Pacific is also seen. The distinct difference of 30-60-day velocity potential between the two types of El Niño winters can be seen in the eastern Indian Ocean and western Pacific (Fig. 3c). The same conclusion that 30-60-day velocity potential is weakened during EP El Niño winters and strengthened during CP El Niño winters, can be derived.

4. Comparison of MJO propagation during the two types of El Niño
  • Generally speaking, MJO-related convection emerges over the tropical western Indian Ocean, then weakens over the Maritime Continent, strengthens again over the western Pacific, and finally quickly dies out over the dateline (e.g., Madden and Julian, 1971; Yuan et al., 2014).

    Figure 4.  Composite anomalies of 30-60-day OLR (units: W m$^-2$) during EP El Niño winters by phase.

    Figure 5.  As in Fig. 4 but for CP El Niño winters (units: W m$^-2$).

    Figure 6.  Composite anomalies of 30-60-day OLR (units: W m$^-2$) in the tropics (averaged over 10$^\circ$S-10$^\circ$N) during (a) EP El Niño winters and (b) CP El Niño winters.

    To understand the influence of the two types of El Niño events on the eastward propagation of MJO-related convection, composite patterns of 30-60-day OLR for each of the eight MJO phases are shown in Figs. 4 and 5, based on the RMM index. The intensity and propagation of the MJO are quite different between the two types of El Niño events. During EP El Niño winters (Fig. 4), the MJO emerges over the tropical eastern Indian Ocean, then develops from the Maritime Continent to the western Pacific, and finally weakens over the central Pacific. The convection can spread to the tropical central Pacific (near 120°W). By contrast, during CP El Niño winters (Fig. 5), the MJO emerges over the western Indian Ocean, and can maintain or even enhance its intensity over the region west of 120°E, which is in agreement with previous studies (Kessler, 2001; Tam and Lau, 2005). However, the propagation tends to be concentrated to the west of the dateline, and convection anomalies become greatly reduced to the east of 180°.

    These eastward-propagation features of the MJO are also illustrated by Fig. 6, which shows the composite tropical (10°S-10°N) OLR anomalies based on the eight phases of the MJO. During EP El Niño winters (Fig. 6a), the MJO occurs near 60°E in phase 2. When the 30-60-day convection moves to 120°E in phase 4, it reaches its strongest intensity. It then weakens in phases 5 and 6, and strengthens again in phase 7. Finally, it dies out near 120°W. During CP El Niño winters, the MJO starts from the west of 60°E in phase 1. It continuously intensifies from phases 2 to 4, and then weakens and maintains its intensity until 180°. Comparing the extent of eastward propagation, the MJO can spread to 120°W during EP El Niño winters and stop propagating near 180° during CP El Niño winters. Another difference between EP and CP El Niño winters is that the MJO during CP El Niño winters tends to have a standing oscillation feature over the eastern Indian Ocean and western Pacific, which is mainly the result of the MJO in 1990 [Fig. S1 in Electronic Supplementary Material (ESM)], when this feature of standing oscillation was predominant.

    Therefore, the above results suggest that the two types of El Niño may have different impacts on the eastward propagation of the MJO. During EP El Niño winters, the abnormally warm sea area is situated in the eastern Pacific, and the MJO can propagate to the eastern Pacific. By contrast, during CP El Niño winters, with the SST positive anomaly moving to the central Pacific, the MJO can only propagate to the dateline.

  • Figure 6 also demonstrates that the phase speed of the MJO displays different features during the two types of El Niño winters. During EP El Niño, the MJO moves slowly in phases 2 and 3 (roughly 0.16° d-1), but rapidly in phases 4 and 5 (0.74° d-1). This means that the phase speed of the MJO is significantly slower over the Indian Ocean and faster over the Maritime Continent during EP El Niño. However, during CP El Niño, the MJO propagates relatively quickly in phases 2 and 3 (0.43° d-1), but slowly in phases 4 and 5 (0.21° d-1). These phase speeds are estimated by the longitudes of MJO propagation and corresponding days

    In addition, we have counted the numbers of MJO days and calculated the proportion in each phase during the two types of El Niño winters, and compared them with normal winters, i.e., the winters of both Niño3 index and EMI anomalies being lower than one standard deviation (Table 1). Overall, the occurrence distributions in different phases is relatively equal, but slightly more frequent in phases 6 and 7 during normal winters, which accounts for 30% of total occurrence. However, it decreases sharply when El Niño occurs, especially during EP events (only 20%). The frequency of MJO occurrence is relatively high in phases 2 and 3 during EP El Niño winters, in which it approaches 35% of the total occurrence. Phases 4 and 5 (only 18%) show the least frequent occurrence. In contrast, the MJO occurs more often in phases 4 and 5 (nearly 40%) during CP El Niño winters, and less frequently in phases 8 and 1 (less than 15%). This suggests that the MJO may occur more frequently over the tropical Indian Ocean during EP El Niño winters, while it may favor the Maritime Continent during CP El Niño winters.

5. Summary and discussion
  • The impacts of two types of El Niño on the MJO during boreal winter are investigated in this study. It is found that the characteristics of MJO activity are quite different between the two types of El Niño.

    The variance of 30-60-day OLR and 200-hPa velocity potential are applied to identify the MJO intensity. Composites of MJO intensity are presented for EP El Niño and CP El Niño, as well as their differences. It is found that the strongest difference occurs between the tropical eastern Indian Ocean and western Pacific. Both variables lead to the conclusion that the intensity of the MJO is weak during EP El Niño, but stronger during CP El Niño.

    Additionally, the propagation features of MJO-related convection are contrasted between the two types of El Niño. The composite of 30-60-day OLR for the eight MJO phases is based on the RMM index. The evolution of the MJO presents the extent of eastward propagation during the different types of El Niño. For EP El Niño, MJO-related convection emerges over the eastern Indian Ocean and can propagate further eastward into the central Pacific (to nearly 120°W). During CP El Niño winters, MJO-related convection emerges over the western Indian Ocean and can only propagate to the dateline, and there are no clear convection anomalies to the east of 180°. The implication of this finding is that the propagation extent of the MJO may be bounded to the abnormally warm area over the tropical Pacific.

    We also find that the frequency in the eight phases of the MJO differs between the two types of El Niño. In general, MJO-related convection appears more frequently over the western Pacific during boreal winter (Lafleur et al., 2015). However, the occurrence of MJO-related convection is relatively high (nearly 35%) over the Indian Ocean (phases 2 and 3), while it is low (only 18%) over the Maritime Continent (phases 4 and 5) during EP El Niño. During CP El Niño winters, the most frequent occurrence lies over the Maritime Continent (nearly 40%) and the least frequent occurrence lies over the Western Hemisphere (less than 15%).

    This study does not investigate why there are differences in MJO intensity, propagation and occurrence against the background of the two types of El Niño. Our hypothesis, however, is that the intensity of the MJO may be linked to the anomalous convection. During EP El Niño winters, there are positive OLR anomalies over the eastern Indian Ocean and western Pacific (Fig. S2 in ESM), which may weaken the 30-60-day convection. During EP El Niño winters, the enhanced convection appears over the tropical eastern Pacific, which supports the eastward propagation of 30-60-day convection. However, during CP El Niño winters, the enhanced convection moves to the central Pacific, which limits the eastward propagation. In addition to these possible effects of convection on MJO intensity and propagation, MJO occurrence may also be affected by convection. (Pohl and Matthews, 2007) hypothesized that the moisture conditions may influence the propagation speed of the MJO. In this study, during EP El Niño, humidity in the lower troposphere is higher over the western Indian Ocean, and lower over the Maritime Continent (Fig. S3 in ESM). According to the hypothesis of (Pohl and Matthews, 2007), these moisture anomalies may induce slower propagation speeds over the Indian Ocean and higher speeds over the Maritime Continent, thus resulting in greater and less occurrence over these two regions, respectively. During CP El Niño, however, the humidity anomalies are much weaker in the Indian Ocean and Maritime Continent, implying that other mechanisms may be at work. In summary, the mechanisms responsible for the differences in MJO intensity, propagation and occurrence between the two types of El Niño should be further studied.

Reference

Catalog

    /

    DownLoad:  Full-Size Img  PowerPoint
    Return
    Return