更全的杂志信息网

Energetic processes regulating the strength of MJO circulation over the Maritime Continent during two types of El Niño

更新时间:2016-07-05

1. Introduction

The MJO (Madden and Julian 1971) is the dominant mode of intraseasonal variability over the tropics. Its convection-circulation coupled signals are of a planetary scale and propagate eastward along the equator. During its journey, the MJO interacts with different weather and climate systems, such as tropical cyclones (Camargo, Wheeler,and Sobel 2009; Maloney and Dickinson 2003), extreme rainfall events (Mao, Sun, and Wu 2010; Yang 2010; Hsu,Lee, and Ha 2016), monsoon activity (Lee et al. 2013), and ENSO (Slingo et al. 1999; Kessler 2001; Hendon, Wheeler,and Zhang 2007), in fluencing the evolution and amplitude of these weather and climate systems. Advancing our understanding of MJO variation at different timescales and mean flow-MJO-eddies scale interactions can bridge the gap between weather forecasting and climate prediction(Waliser 2006), which is key to developing a seamless prediction system (Palmer et al. 2008).

淼哥笑了笑:“理论是个好东西,既然你这么喜欢理论,我来给你算算。一般小区里的游泳池50米长、10米宽、1.5~2米深,那大概是750~1000立方米水。正常精液的标准为:精液量2~6毫升,精子数大于2000万/毫升。正常人精液中精子数为4000~6000万/毫升,当低于1000万/毫升时,就有可能需要借助辅助生育技术。

The interannual variation of MJO activity is generally linked to the changes in equatorial SST at the interannual timescale. Earlier studies (Slingo et al. 1999; Kessler 2001;Hendon, Wheeler, and Zhang 2007) focused on the interaction between the MJO and conventional ENSO, which has the maximum SST warming over the eastern equatorial Pacific (referred to as eastern Pacific El Niño, or EP El Niño).Accompanied by the eastward shift of anomalous SST warming, the strengthened MJO signals tend to extend farther east toward the eastern Pacific, while weakened MJO activity appears over the western Pacific, during EP El Niño years (Hendon, Zhang, and Glick 1999; Kessler 2001). Since the beginning of the twenty- first century, a new type of El Niño with a significant SST warming over the central Pacific(referred to as CP El Niño) has been observed frequently.CP El Niño induces a different modulating effect on MJO activity compared to EP El Niño (Gushchina and Dewitte 2012; Feng et al. 2015; Yuan, Li, and Ling 2015; Chen, Ling,and Li 2016; Hsu and Xiao 2017). The suppressed MJO over the western Pacific and Maritime Continent during CP El Niño events is of less significance than that during EP El Niño years. Some studies have suggested that the difference in the MJO over the western Pacific/Maritime Continent could be related to the suppressed effect of the Walker circulation and the anticyclonic circulation anomaly near the Philippine Sea during different types of El Niño events (Feng et al. 2015; Yuan, Li, and Ling 2015; Chen,Ling, and Li 2016).

(5)K+Na:1~3质量及分子25.0,26.2,18.0;4~6质量及分子20.6,23.3,16.1;7~8质量及分子40.0,9.6。

Although possible relationships between the anomalous mean flow and MJO associated with each type of El Niño have been indicated (Feng et al. 2015; Yuan, Li, and Ling 2015; Chen, Ling, and Li 2016), the detailed processes and relative contributions of these anomalous circulations to the changes in the MJO need to be investigated. In addition to large-scale circulations, synoptic-scale variability is also vigorous over the Maritime Continent (Chang, Harr,and Chen 2005) and may have in fluences on MJO activity through upscaled feedback. In this study, we examine the main source of increased MJO kinetic energy (KE) during CP El Niño events based on a newly proposed MJO KE budget equation derivation. We also quantitatively discuss how scale interactions among low-frequency background flow, the MJO, and high-frequency disturbances modulate the MJO strength during different types of El Niño.

In this study, we derive a new MJO KE budget equation, in which the low-frequency background mean flow-MJO interaction and high-frequency disturbances-MJO interaction are formulated, to quantitatively examine the physical processes modulating the MJO activity during different El Niño events. The results show that both the barotropic energy conversion from background mean flow to MJO KE (CKL-M) and baroclinic energy conversion from MJO available potential energy to KE (CE) contribute positively to the enhanced MJO KE over the Maritime Continent during the mature phase of CP El Niño. Among the three-dimensional large-scale circulation anomalies during CP El Niño, the low-level convergence and cyclonic anomalies related to a weakened descending branch (or an upward anomaly) of the Walker circulation over the Maritime Continent and the reduced Philippine anticyclonic anomaly play crucial roles in favoring the KE conversion from mean flow to MJO. Different from the positive contribution of CKL-M occurring at the mid-to-lower troposphere, the enhanced CE is the major contributor to the strengthened MJO KE at the upper troposphere during CP El Niño. Based on the diagnosis of interaction between MJO and high-frequency disturbances (CKH-M), we find that the high-frequency variability over the Maritime Continent is enhanced during CP El Niño because it obtains more KE from the MJO.

2. Methodology

To identify the deep convection associated with MJO and high-frequency disturbances, we use daily OLR on a 2.5 × 2.5° grid from NOAA (Liebmann 1996). The changes in dynamic and thermodynamic conditions and their in fluences on the MJO are examined using daily-averaged fields of horizontal wind (u, v), vertical p-velocity (ω), temperature (T), and geopotential (ϕ) from ERA-Interim (Dee et al. 2011). These fields have a resolution of 1.5 × 1.5 and 19 levels from 1000 to 100 hPa with 50-hPa intervals. All data cover the period 1979–2014.

To reduce data uncertainty, the average of monthly SST data derived from HadISST1 (Rayner et al. 2003) and ERSST.v3 (Smith et al. 2008) are used to categorize the two different types of El Niño. Using the method proposed by Yeh et al.(2009) and others (e.g. McPhaden, Lee, and McClurg 2011;Ren and Jin 2011), the CP and EP cases are classified based on the amplitude of SST anomalies over the Niño3 (5°S–5°N,90°–150°W) region and the Niño4 (5°S–5°N, 160°E–150°W)region. Those El Niño events with a larger Niño3 than Niño4(larger Niño4 than Niño3) warming during boreal winter are considered as EP El Niño (CP El Niño) events. Four EP El Niño events (1982/83, 1986/87, 1991/92, and 1997/98) and five CP El Niño events (1994/95, 2002/03, 2004/05, 2006/07,and 2009/10) are selected for the study period.

To quantitatively examine the physical processes modulating the MJO activity during the two types of El Niño,the MJO KE budget is derived and then diagnosed. The conventional energy budget equation is derived by partitioning the meteorological variable into two parts –background mean flow and perturbation – to diagnose the energy conversion between the two (Oort 1964; Lau and Lau 1992; Maloney and Dickinson 2003; Hsu and Chih-Hua 2009). To understand the scale interactions of MJO with both the mean flow and high-frequency disturbances, we decompose a variable into three parts in the time domain, including the low-frequency mean flow (>90 days), MJO (20–90 days), and high-frequency disturbances (< 20 days), as follows:

wherecates the horizontal wind vector. ∇ and ∇3 indicate the twoand three-dimensional gradient operators, respectively. R is the gas constant. P is pressure. D includes dissipation and subgrid-scale effects. The terms on the rhs of Equation(2) represent the processes that modulate the tendency of MJO KE. The first two terms, CKL-M and CKH-M, indicate the barotropic energy conversion from the low-frequency mean flow and high-frequency disturbances to MJO,respectively. Term CE, the baroclinic conversion from the MJO available potential energy to MJO KE, is related to the interaction between MJO circulation and convection.When the MJO upward (downward) motion occurs over the convectively heating (cooling) area, the MJO gains KE through the baroclinic energy conversion. The advection of MJO induced by the three-dimensional wind fields is denoted by BK. The convergence of MJO geopotential flux (Bϕ) can also induce MJO KE. Note that only terms CK(CKL-M and CKH-M) and CE are related to the major sources of MJO KE via generation and conversion processes, while the other terms (BK and Bϕ) represent redistributions of MJO KE.

The distinct impacts of CP and EP El Niño events on western Pacific MJO activity have been documented previously(Gushchina and Dewitte 2012; Feng et al. 2015; Yuan, Li,and Ling 2015; Chen, Ling, and Li 2016; Hsu and Xiao 2017).However, the physical mechanisms responsible for the differences in MJO associated with the two types of El Niño have not been fully understood. Particularly, the western Pacific/Maritime Continent is the region that undergoes vigorous multi-scale variability. How and to what extent the changes in background mean flow and high-frequency disturbances in fluence the MJO during the CP and EP El Niño need further elucidation.

创建中国—东盟知识产权港,以此集聚创新资源、吸引高价值知识产权特别是专利技术落地实施、交易流转和转移转化,为我国与东盟国家开展国际产能合作项目、提高与沿线国家贸易额以及供给侧结构性改革和产业转型升级提供有力支撑。最根本的是,通过知识产权港的建设,汇聚外部先进创新要素和技术资源,以外源性技术动力推动广西产业升级,壮大广西产业规模,并由此逐步培育和形成一批运营规模较大、创新能力强、市场竞争能力较为充分的区域性国际跨国公司,从源头提升广西对外合作的内生能力。

where the overbar denotes the low-frequency component;the prime and asterisk denote the 20–90-day MJO and the high-frequency disturbances of shorter than 20 days,respectively. Here, we apply Lanczos band-pass filtering(Duchon 1979), 90-day low-pass and 20-day high-pass filtering to extract the MJO, low-frequency mean flow and high-frequency disturbances, respectively. This derivation strategy is similar to that used in Hsu, Li, and Tson (2011)and Tsou, Hsu, and Hsu (2014), although these studies focused on the KE sources of synoptic-scale eddies rather than the MJO.

The results of Figures 1 and 2 are based on the seasonal average. To ensure the timing when the two types of El Niño exert differential in fluences on the MJO circulation, we analyze the temporal evolution of equatorial MJO KE (not shown). The significant increase in MJO KE during CP El Niño appears over the Maritime Continent from mid-September of the El Niño developing year to the following February. This con firms our composite analysis for the period of autumn-winter (September-February) is reasonable.

3. Results

Figures 1 and 2 show the changes in subseasonal variability (including 20–90-day MJO and high-frequency disturbances of less than 20 days), large-scale circulations, and SST patterns during the mature phase (autumn-winter)of the two types of El Niño events relative to their climatological conditions. When the SST warming maximizes at the eastern equatorial Pacific (Figure 1(a)), the ascending motion of the Walker circulation and trade winds are reduced significantly. A low-level divergence (Figure 1(a))associated with a downward anomaly of the Walker circulation (Figure 2(a)) can be observed over the Maritime Continent. An anticyclonic anomaly appears near the Philippine Sea (Wang, Wu, and Fu 2000). Along with the changes in background dynamic and thermodynamic conditions, both the atmospheric MJO (Figure 1(a)) and high-frequency eddies (Figure 2(a)) are weakened over the western Pacific and Maritime Continent during the EP El Niño events.

Figure 1. Kinetic energy of 20–90-day MJO (shading; units: m2 s−2), monthly SST (contours; units: °C), and monthly 850-hPa wind (vectors;units: m s−1) anomalies during autumn-winter (September-February) of (a) EP and (b) CP El Niño events relative to their climatological states. The differences in these fields between CP and EP El Niño events (CP minus EP) are shown in (c). Stippling marks the regions with the changes in MJO KE exceeding the 90% con fidence level.

As the warm SST anomaly shifts westward during CP El Niño events, the low-level divergence and anticyclonic anomalies associated with the descending anomaly of the Walker circulation are of less significance (Figures 1(b)and 2(b)), compared to those during EP El Niño events(Figures 1(a) and 2(a)). Thus, the differences in large-scale circulations between CP and EP El Niño events exhibit lowlevel convergence and cyclonic anomalies (Figure 1(c))associated with ascending anomalies over the Maritime Continent (Figure 2(c)). The amplitude of transient eddies at the intraseasonal and synoptic timescales vary obviously over the Maritime Continent during the two types of El Niño. The MJO has higher KE over the Maritime Continent during CP El Niño than during EP El Niño (Figure 1(b)and (c)), consistent with the findings of Chen, Ling, and Li (2016). Although the high-frequency variability tends to weaken over the Maritime Continent during both CP and EP El Niño events relative to the climatology (Figure 2(a) and (b)), the reduction in KE of high-frequency disturbances is less during CP El Niño events compared to that during EP El Niño events (Figure 2(c)).

许多初中英语阅读教学模式以纯理论式和灌输式教学为主,过于重视学生英语考试成绩的提升,忽略了学生英语素养以及英语运用能力的培养,此外,大多数初中英语老师采用统一化的课堂教学模式,对班级内所有学生进行无差别教学,忽略了学生的个体差异性,不仅造成了部分英语水平较高学生课堂学习时间的浪费,也增加了一些英语基础薄弱学生学习英语知识和紧跟老师授课节奏的难度,不利于班级整体英语水平的提升。

To compare the amplitude and life cycle of MJO KE over the Maritime Continent between the two types of El Niño, enhanced MJO circulation events occurring over the region (10°S–10°N, 90°–150°E) with significant changes in MJO KE between the CP and EP El Niño events (Figure 1(c)) are selected and composited. An enhanced MJO circulation event is identified by the MJO KE over the key region of (10°S–10°N, 90°–150°E) exceeding one standard deviation. The date with maximum KE for each enhanced MJO event is de fined as Day 0. Figure 3(a) compares the KE evolution of composited active MJO events in CP and EP El Niño years. The enhanced MJO KE is obvious over the Maritime Continent during CP El Niño and larger than that during EP El Niño, in agreement with the result in Figure 1(c). From nine days before the MJO KE reaches its maximum, a positive tendency of MJO KE can be found for both CP and EP El Niño (not shown). The growth rate of MJO KE is larger during CP El Niño than during EP El Niño,accounting for the enhanced MJO KE over the Maritime Continent during CP El Niño (Figure 3(a)).

Figure 2. Kinetic energy of less than 20-day eddies (shading; units: m2 s−2) and 500-hPa monthly vertical p-velocity (contours; units:Pa s−1) of (a) EP and (b) CP El Niño events relative to their climatological states. The differences in these fields between CP and EP El Niño events (CP minus EP) are shown in (c). Stippling marks the regions with the changes in eddy KE exceeding the 90% con fidence level. The thick solid and dashed contours denote the anomalies of 500-hPa vertical p-velocity at 0.2 and −0.2 Pa s−1, respectively.

To understand the key processes modulating the sources of MJO KE associated with the two types of El Niño,the column (1000–200 hPa) MJO KE budget (Equation (2))is diagnosed during Day −9 to 0 (Figure 3(b)). The larger growth rate of MJO KE during CP El Niño is mainly from the barotropic energy conversion from mean flow to MJO (CKL-M) and the baroclinic energy conversion from the MJO available potential energy to KE (CE). The positive contribution of CKL-M via scale interaction between anomalous mean flow and MJO appears in the mid-to-lower troposphere(Figure 4(a)). However, the CE associated with the MJO circulation-convection coupled feedback plays an important role in maintaining the increased MJO KE during CP El Niño at the upper troposphere (Figure 4(b)). Although the MJO obtains more KE during CP El Niño, it provides more KE to high-frequency eddies through CKH-M (Figure 3(b)) in the meantime, supporting the enhanced high-frequency variability over the Maritime Continent (Figure 2(c)). The redistributions of MJO KE associated with the advection process and geopotential flux contribute negatively to the increased MJO KE during CP El Niño (Figure 3(b)).

Figure 3 (a) Differences in composited MJO kinetic energy (blue curve; units: m2 s−2) and its tendency (red curve; 10−6 m2 s−3) over the Maritime Continent (10°S−10°N, 90°−150°E) between CP and EP El Niño events. Day 0 represents the date when the MJO kinetic energy reaches its maximum for each event. (b) MJO kinetic energy budget terms integrated between 1000 and 200 hPa over the Maritime Continent during Day −9 to 0 for EP (blue bars) and CP (red bars) El Niño events and their differences (black bars). Units: kg s−3. (c) As in(b) except for individual terms of CKL-M.

The individual terms of CKL-M are then compared to identify the major contributors (Figure 3(c)). The MJO eddy momentum working on background zonal flow conto zonal wind convergence and cyclonic anomalies show a large contribution at the lower troposphere, while the third term associated with vertical wind shear maximizes around 400 hPa where the zonal wind changes its sign(not shown). During CP El Niño, the descending anomaly of the Walker circulation and low-level divergence anomaly over the Maritime Continent are weaker compared to those during EP El Niño (Figure 2(a) and (b)). Meanwhile,the Philippine Sea anticyclonic anomaly tends to be weakened during CP El Niño (Figure 1(c)), similar to the findings of Yuan, Yang, and Zhang (2012). These large-scale anomalies associated with the westward shift of the equatorial SST warming pattern (i.e. CP El Niño) generate enhanced barotropic energy conversion from the mean flow to MJO KE as they work with MJO eddy momentum fluxes.

Understanding the multi-scale interaction is a key step for developing seamless prediction (Waliser 2006; Palmer et al. 2008), while the quantitative diagnosis of scale interactions is still challenging. The new MJO KE budget equation proposed in this study can help to diagnose how and to what extent the MJO interacts with the mean flow and with the high-frequency disturbances. The energy source of the MJO is also examined quantitatively using the MJO KE budget equation. Specifically, we use this diagnostic approach to explain the modulation of the MJO by different types of El Niño. We plan to carry out more studies related to MJO dynamics and scale interactions based on the diagnosis of the MJO KE budget equation.

Figure 4. Vertical pro files of (a) CKL-M and (b) CE over the Maritime Continent (10°S–10°N, 90°–150°E) averaged over the period of Day−9 to 0, when the MJO kinetic energy shows positive tendency in Figure 3, for MJO cases during CP El Niño (red curve), EP El Niño (blue curve), and their differences (black curve). Units: 10−5 m2 s−3.

4. Summary

By multiplying u′ and v′ on both sides of the MJO-filtered horizontal momentum equations, respectively,the MJO KE budget equation is obtained as

定义7 称ftr/tr-1(A(tr)/A(tr-1)),(2≤r≤m)为马尔可夫过程{ξ(t),t∈T},(t1

图7展示了FISE的查找流程。当一个报文到达时,FISE首先在TCAM中分别匹配目的前缀和源前缀,通过目的表和源表可得指向TD单元的目的索引号和源索引号,最后利用TD单元存储的下一跳索引号,在映射表中查找到下一跳的信息。

Acknowledgments

The authors would like to thank the anonymous reviewers for their help in improving the manuscript.

Disclosure statement

No potential conflict of interest was reported by the authors.

Funding

This study was supported by the National Natural Science Foundation of China [grant number 41375100]; the National Basic Research Program of China [973 Program, grant number 2015CB453200]; and the Natural Science Foundation of Jiangsu Province [grant number BK20140046].

References

Camargo, S. J., M. C. Wheeler, and A. H. Sobel. 2009. “Diagnosis of the MJO modulation of tropical cyclogenesis using an empirical index.” Journal of the Atmospheric Sciences 66 (10):3061–3074.

Chang, C.-P., P. A. Harr, and H. J. Chen. 2005. “Synoptic disturbances over the equatorial South China Sea and western Maritime Continent during boreal winter.” Monthly Weather Review 133 (3): 489–503.

Chen, X., J. Ling, and C. Y. Li. 2016. “Evolution of the Madden–Julian oscillation in two types of El Niño.” Journal of Climate 29 (5): 1919–1934.

Dee, D. P., S. M. Uppala, A. J. Simmons, P. Berrisford, P. Poli, S.Kobayashi, U. Andrae, et al. 2011. “The ERA-Interim reanalysis:con figuration and performance of the data assimilation system.” Quarterly Journal of the Royal Meteorological Society 137 (656): 553–597.

Duchon, C. E. 1979. “Lanczos filtering in one and two dimensions.”Journal of Applied Meteorology 18 (8): 1016–1022.

Feng, J., P. Liu, W. Chen, and X. C. Wang. 2015. “Contrasting Madden–Julian Oscillation activity during various stages of EP and CP El Niños.” Atmospheric Science Letters 16 (1): 32–37.

Gushchina, D., and B. Dewitte. 2012. “Intraseasonal tropical atmospheric variability associated with the two flavors of El Niño.” Monthly Weather Review 140 (11): 3669–3681.

Hendon, H. H., C. Zhang, and J. D. Glick. 1999. “Interannual variation of the Madden–Julian oscillation during austral summer.” Journal of Climate 12 (8): 2538–2550.

Hendon, H. H., M. C. Wheeler, and C. Zhang. 2007. “Seasonal dependence of the MJO–ENSO relationship.” Journal of Climate 20: 531–543.

Hsu, P.-C., and C.-H. Tsou, 2009. “Eddy energy along the tropical storm track in association with ENSO.” Journal of the Meteorological Society of Japan Ser. II 87(4): 687–704.

Hsu, P.-C., and T. Xiao. 2017. “Differences in the initiation and development of the Madden–Julian Oscillation over the Indian Ocean associated with two types of El Niño.” Journal of Climate 30 (4): 1397–1415.

Hsu, P. C., T. Li, and C. H. Tsou. 2011. “Interactions between Boreal Summer Intraseasonal Oscillations and Synoptic-Scale Disturbances over the Western North Pacific. Part I: Energetics Diagnosis*.” Journal of Climate 24 (3): 927–941.

Hsu, P.-C., J.-Y. Lee, and K.-J. Ha. 2016. “In fluence of boreal summer intraseasonal oscillation on rainfall extremes in southern China.” International Journal of Climatology 36 (3):1403–1412.

Kessler, W. S. 2001. “EOF representations of the Madden–Julian oscillation and its connection with ENSO.” Journal of Climate 14 (13): 3055–3061.

Lau, K.-H., and N.-C. Lau. 1992. “The energetics and propagation dynamics of tropical summertime synoptic-scale disturbances.” Monthly Weather Review 120 (11): 2523–2539.

Lee, J.-Y., B. Wang, M. C. Wheeler, X. H. Fu, E. Duane, and W.-S.Kang. 2013. “Real-time multivariate indices for the boreal summer intraseasonal oscillation over the Asian summer monsoon region.” Climate Dynamics 40 (1-2): 493–509.

Liebmann, B., and C. A. Smith 1996. “Description of a complete(interpolated) outgoing longwave radiation dataset.” Bulletin of the American Meteorological Society 77: 1275–1277.

Madden, R. A., and P. R. Julian. 1971. “Detection of a 40–50 day oscillation in the zonal wind in the tropical Pacific.” Journal of the Atmospheric Sciences 28 (5): 702–708.

Maloney, E. D., and M. J. Dickinson. 2003. “The intraseasonal oscillation and the energetics of summertime tropical western North Pacific synoptic-scale disturbances.” Journal of the Atmospheric Sciences 60 (17): 2153–2168.

Mao, J. Y., Z. Sun, and G. X. Wu. 2010. “20–50-day oscillation of summer Yangtze rainfall in response to intraseasonal variations in the subtropical high over the western North Pacific and South China Sea.” Climate Dynamics 34 (5): 747–761.

McPhaden, M. J., T. Lee, and D. McClurg. 2011. “El Niño and its relationship to changing background conditions in the tropical Pacific Ocean.” Geophysical Research Letters 38 (15):L15790.

Oort, A. H. 1964. “On estimates of the atmospheric energy cycle.”Monthly Weather Review 92 (11): 483–493.

Palmer, T. N., F. J. Doblas-Reyes, A. Weisheimer, and M. J. Rodwell.2008. “Toward seamless prediction: Calibration of climate change projections using seasonal forecasts.” Bulletin of the American Meteorological Society 89 (4): 459–470.

Rayner, N. A., D. E. Parker, E. B. Horton, C. K. Folland, L. V.Alexander, P. Rowell, C. Kent, and A. Kaplan. 2003. “Global analyses of sea surface temperature, sea ice, and night marine air temperature since the late nineteenth century.”Journal of Geophysical Research: Atmospheres 108 (D14): 4407.doi: 10.1029/2002JD002670.

Ren, H.-L., and F.-F. Jin. 2011. “Niño indices for two types of ENSO.” Geophysical Research Letters 38 (4): L04704.

Slingo, J. M., D. P. Rowe, R. Sperber, and F. Nortley. 1999. “On the predictability of the interannual behaviour of the Madden-Julian Oscillation and its relationship with El Niño.” Quarterly Journal of the Royal Meteorological Society 125 (554): 583–609.

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).” Journal of Climate 21: 2283–2296.

Tsou, C.-H., H.-H. Hsu, and P.-C. Hsu. 2014. “The role of multiscale interaction in synoptic-scale eddy kinetic energy over the western North Pacific in autumn.” Journal of Climate 27 (10):3750–3766.

Waliser, D. E. 2006. “Intraseasonal variability.” The Asian Monsoon.Springer Berlin Heidelberg, 203–257.

Wang, B., R. Wu, and X. Fu. 2000. “Pacific–East Asian teleconnection: How does ENSO affect East Asian climate?”Journal of Climate 13 (9): 1517–1536.

Yang, J. 2010. “Biweekly and 21–30-day variations of the subtropical summer monsoon rainfall over the lower reach of the Yangtze River basin.” Journal of Climate 23 (5): 1146–1159.

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 (7263):511–514.

Yuan, Y., S. Yang, and Z. Zhang. 2012. “Different evolutions of the Philippine Sea anticyclone between the eastern and central Pacific El Niño: Possible effects of Indian Ocean SST.” Journal of Climate 25 (22): 7867–7883.

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.

HSU Pang-Chi,FU Zhen,XIAO Ting
《Atmospheric and Oceanic Science Letters》2018年第2期文献
Preface 作者:Hui-Jun Wang,Ola M.Johannessen

服务严谨可靠 7×14小时在线支持 支持宝特邀商家 不满意退款

本站非杂志社官网,上千家国家级期刊、省级期刊、北大核心、南大核心、专业的职称论文发表网站。
职称论文发表、杂志论文发表、期刊征稿、期刊投稿,论文发表指导正规机构。是您首选最可靠,最快速的期刊论文发表网站。
免责声明:本网站部分资源、信息来源于网络,完全免费共享,仅供学习和研究使用,版权和著作权归原作者所有
如有不愿意被转载的情况,请通知我们删除已转载的信息