Article Contents

Leading Pattern of Spring Drought Variability over East Asia and Associated Drivers

Funding:

National Natural Science Foundation of China 42230603

National Natural Science Foundation of China 42275020

Guangdong Major Project of Basic and Applied Basic Research 2020B0301030004

Guangdong Province Key Laboratory for Climate Change and Natural Disaster Studies 2020B1212060025

Innovation Group Project of Southern Marine Science and Engineering Guangdong Laboratory (Zhuhai) 311021001

Open Fund of State Key Laboratory of Satellite Ocean Environment Dynamics, Second Institute of Oceanography, MNR QNHX2310


doi: 10.3724/j.1006-8775.2024.001

  • Drought events have become more frequent and intense over East Asia in recent decades, leading to huge socioeconomic impacts. Although the droughts have been studied extensively by cases or for individual regions, their leading variability and associated causes remain unclear. Based on the Standardized Precipitation Evapotranspiration Index (SPEI) and ERA5 reanalysis product from 1979 to 2020, this study evealuates the severity of spring droughts in East Asia and investigates their variations and associated drivers. The results indicate that North China and Mongolia have experienced remarkable trends toward dryness during spring in recent decades, while southwestern China has witnessed an opposite trend toward wetness. The first Empirical Orthogonal Function mode of SPEI variability reveals a similar seesawing pattern, with more severe dryness in northwestern China, Mongolia, North China, South Korea, and Japan but increased wetness in Southwestern China and southeast Asia. Further investigation reveals that the anomalously dry (wet) surface in North (Southwestern) China is significantly associated with anomalously high (low) temperature, less (more) precipitation, and reduced (increased) soil moisture during the previous winter and early spring, regulated by an anomalous anticyclone (cyclone) and thus reduced (increased) water vapor convergence. The spring dry-wet pattern in East Asia is also linked to cold sea surface temperature anomalies in the central-eastern Pacific. The findings of this study have important implications for improving the prediction of spring drought events in East Asia.
  • 加载中
  • Figure 1.  (a) Climatology, (b) linear trend, and (c) standardized deviation of the spring SPEI over East Asia.

    Figure 2.  (a) Spatial pattern and (b) primary component of EOF1 of variability in the spring SPEI over East Asia. In (b), the black curve and dashed line indicate the original PC1 and its regression trends, respectively, while the bars represent the PC1 with the trends removed.

    Figure 3.  Composite patterns of spring SPEI are based on the years when the standardized deviations of PC1 are (a) more and (b) less than 1.5. (c) Difference between (a) and (b).

    Figure 4.  Lag-lead correlations between SPEIs in different lags and the detrend PC1 during 1979–2020.

    Figure 5.  Lag-lead correlations of 2-m temperature (left), precipitation (middle), and soil moisture (right) in different leads with the detrend PC1.

    Figure 6.  Lagged regressions of geopotential height (units: dagpm) and horizontal winds (units: m s–1) in different leads at 850 hPa (left), 500 hPa (middle), and 300 hPa (right) onto the detrend PC1. Areas with significant anomalies exceeding the 95% confidence level are stippled.

    Figure 7.  Regressions of the vertical integrals of (a) water vapor flux (vectors, units: kg m–1 s–1) and divergence (units: kg m–2 s–1), (b) total cloud cover, and (c) solar download radiation in spring onto the detrend PC1. Areas with significant anomalies exceeding the 95% confidence level are stippled.

    Figure 8.  (a) Regression between SST in spring and the detrend PC1. (b) Regression of 300-hPa geopotential height and horizontal winds onto the detrend PC1.

    Figure 9.  (a) SST index averaged over the area in the box shown in Fig. 8a (bars) and the detrend PC1 (curve). (b), (c), and (d) respectively represent the regressions of 300-hPa geopotential height and horizontal winds, the water vapor flux and divergence, and the SPEI onto the SST index.

  • [1] CHEN W H, JIA S G, WANG X M, et al. Precipitation trend increases the contribution of dry reduced nitrogen deposition [J]. npj Climate and Atmospheric Science, 2023, 6(1): 62, https://doi.org/10.1038/s41612-023-00390-7
    [2] LI M G, ZHAO Y, LI Y, et al. Interdecadal variation and possible causes of summer extreme precipitation over northern Xinjiang province, northwestern China [J]. Environmental Research Communications, 2023, 5(8): 085001, https://doi.org/10.1088/2515-7620/acec38
    [3] HUANG L, DU H, DANG Y, et al. More frequent consecutive extreme precipitation in the dry regions of China [J]. International Journal of Climatology, 2022, 42(16): 9583–9594, https://doi.org/10.1002/joc.7848
    [4] CHEN H J, WANG S, ZHU J X, et al. Projected changes in the pattern of spatially compounding drought and pluvial events over eastern China under a warming climate [J]. Earth's Future, 2023, 11(5): e2022EF003397, https://doi.org/10.1029/2022EF003397
    [5] KOSTOPOULOU E, GIANNAKOPOULOS C. Projected changes in extreme wet and dry conditions in Greece [J]. Climate, 2023, 11(3): 49, https://doi.org/10.3390/cli11030049
    [6] HAN X, LI Y, YU W D, et al. Attribution of the extreme drought in the horn of Africa during short-rains of 2016 and long-rains of 2017 [J]. Water, 2022, 14(3): 1–16, https://doi.org/10.3390/w14030409
    [7] SANTINI M, NOCE S, ANTONELLI M, et al. Complex drought patterns robustly explain global yield loss for major crops[J]. Scientific Reports, 2022, 12(1): 10.1038, https://doi.org/10.1038/s41598-022-09611-0
    [8] HIGGINS P A, PALMER J G, ANDERSEN M S, et al. Extreme events in the multi-proxy South Pacific drought atlas[J]. Climatic Change, 2023, 176(8): 10.1007, https://doi.org/10.1007/s10584-023-03585-2
    [9] PURICH A, BOSCHAT G, LIGUORI G. Assessing the impact of suppressing Southern Ocean SST variability in a coupled climate model [J]. Scientific Reports, 2021, 11(1): 22069, https://doi.org/10.1038/s41598-021-01306-2
    [10] WANG R, LÜ G N, NING L, et al. Likelihood of compound dry and hot extremes increased with stronger dependence during warm seasons [J]. Atmospheric Research, 2021, 260(24): 105692, https://doi.org/10.1016/j.atmosres.2021.105692
    [11] HEINO M, KINNUNEN P, ANDERSON W, et al. Increased probability of hot and dry weather extremes during the growing season threatens global crop yields [J]. Scientific Reports, 2023, 13(1): 10.1038, https://doi.org/10.1038/s41598-023-29378-2
    [12] SUSKA K, PIOTR P. Application of the Mesohabitat Simulation System (MesoHABSIM) for assessing impact of river [J]. Water, 2020, 12(12): 3356, https://doi.org/10.3390/w12123356
    [13] LI Q Q, CAO Y P, MIAO S L, et al. Spatiotemporal characteristics of drought and wet events and their impacts on agriculture in the Yellow River basin [J]. Land, 2022, 11(4): 556, https://doi.org/10.3390/land11040556
    [14] DONG N, XU X D, CAI W Y, et al. The response of anomalous vertically integrated moisture flux patterns related to drought and flood in southern China to sea surface temperature anomaly [J]. Journal of Tropical Meteorology, 2023, 29(2): 179–190, https://doi.org/10.46267/j.1006-8775.2023.014
    [15] ZHOU K, WANG Y M. Temporal and spatial distribution of multi-scale drought in North China [J]. IOP Conference Series: Earth and Environmental Science, 2020, 428(1): 012057, https://doi.org/10.1088/1755-1315/428/1/012057
    [16] DENG Y, GAO L L, GOU X H. Spatiotemporal drought variation in midlatitude East Asia over the past half millennium [J]. Journal of Geophysical Research: Atmospheres, 2023, 128(6): e2022JD037793, https://doi.org/10.1029/2022JD037793
    [17] ZHANG L, XIAO J F, LI J, et al. The 2010 spring drought reduced primary productivity in southwestern China [J]. Environmental Research Letters, 2012, 7(4): 045706, https://doi.org/10.1088/1748-9326/7/4/045706
    [18] LIU Y Y, HU Z Z, WU R G, et al. Causes and predictability of the 2021 spring southwestern China severe drought [J]. Advances in Atmospheric Sciences, 2022, 39(10): 1766–1776, https://doi.org/10.1007/s00376-022-1428-4
    [19] MALLAPATY S. China's extreme weather challenges scientists studying it [J]. Nature, 2022, 609(7929): 888, https://doi.org/10.1038/d41586-022-02954-8
    [20] QIAO Y, XU W, MENG C N, et al. Increasingly dry/wet abrupt alternation events in a warmer world: observed evidence from China during 1980–2019 [J]. International Journal of Climatology, 2022, 42(12): 6429–6440, https://doi.org/10.1002/joc.7598
    [21] TANG Z F, YANG T, LIN X, et al. Future changes in the risk of compound hot and dry events over China estimated with two large ensembles [J]. PLoS One, 2022, 17(3): e0264980, https://doi.org/10.1371/journal.pone.0264980
    [22] ULLAH I, MA X, ASFAW T G, et al. Projected changes in increased drought risks over South Asia under a warmer climate [J]. Earth's Future, 2022, 10(10): e2022EF002830, https://doi.org/10.1029/2022EF002830
    [23] YU S Q, SHI X H, LIN X C. Interannual variation of East Asian summer monsoon and its impacts on general circulation and precipitation [J]. Journal of Geographical Sciences, 2009, 19(1): 67–80, https://doi.org/10.1007/s11442-009-0067-3
    [24] CEN S X, GONG Y F, LAI X, et al. The relationship between the atmospheric heating source/sink anomalies of Asian monsoon and flood/drought in the Yangtze River basin in the Meiyu period [J]. Journal of Tropical Meteorology, 2015, 21(4): 352–360, https://doi.org/10.16555/j.1006-8775.2015.04.004
    [25] HUANG R H, LIU Y, DU Z C, et al. Differences and links between the East Asian and South Asian summer monsoon systems: characteristics and variability [J]. Advances in Atmospheric Sciences, 2017, 34(10): 1204–1218, https://doi.org/10.1007/s00376-017-7008-3
    [26] WANG H J. The weakening of the Asian monsoon circulation after the end of 1970's [J]. Advances in Atmospheric Sciences, 2001, 18(3): 376–386, https://doi.org/10.1007/BF02919316
    [27] WANG Y F, WANG B, OH J H. Impact of the preceding El Niño on the East Asian summer atmosphere circulation [J]. Journal of the Meteorological Society of Japan, 2001, 79(1B): 575–588, https://doi.org/10.2151/jmsj.79.575
    [28] ZHOU T J, YU R C. Atmospheric water vapor transport associated with typical anomalous summer rainfall patterns in China [J]. Journal of Geophysical Research: Atmospheres, 2005, 110(D8): D08104, https://doi.org/10.1029/2004JD005413
    [29] LI H M, DAI A G, ZHOU T J, et al. Responses of East Asian summer monsoon to historical SST and atmospheric forcing during 1950-2000 [J]. Climate Dynamics, 2010, 34(4): 501-514, https://doi.org/10.1007/s00382-008-0482-7
    [30] ZHOU X J, ZHAO P, LIU G, et al. Characteristics of decadal-centennial-scale changes in East Asian summer monsoon circulation and precipitation during the Medieval Warm Period and Little Ice Age and in the present day [J]. Chinese Science Bulletin, 2011, 56(28): 3003, https://doi.org/10.1007/s11434-011-4651-4
    [31] ZHANG H L, ZHANG Q, YUE P, et al. Aridity over a semiarid zone in northern China and responses to the East Asian summer monsoon [J]. Journal of Geophysical Research: Atmospheres, 2016, 121(23): 13,901-13,918, https://doi.org/10.1002/2016JD025261
    [32] SIMMONDS I, BI D, HOPE P. Atmospheric water vapor flux and its association with rainfall over China in summer [J]. Journal of Climate, 1999, 12(5): 1353–1367, https://doi.org/10.1175/1520-0442(1999)012<1353:AWVFAI>2.0.CO;2 doi:
    [33] CAO Q, HAO Z C, ZHOU J W, et al. Impacts of various types of El Niño-Southern Oscillation (ENSO) and ENSO Modoki on the rainy season over the Huaihe River basin [J]. International Journal of Climatology, 2019, 39(5): 2811–2824, https://doi.org/10.1002/joc.6002
    [34] WANG Q, LI S L. Different summer rainfall anomaly patterns in northeast China associated with two kinds of El Niño events [J]. Journal of Tropical Meteorology, 2020, 26(2): 223-230, https://doi.org/10.46267/j.1006-8775.2020.020
    [35] WANG W Z, TANG H Y, LI J P, et al. Spatial-temporal variations of extreme precipitation characteristics and its correlation with El Niño-Southern Oscillation during 1960–2019 in Hubei Province, China [J]. Atmosphere, 2022, 13(11): 1922, https://doi.org/10.3390/atmos13111922
    [36] KIEH J T, TRENBERTH K E. Earth's annual global energy budget [J]. Bulletin of the American Meteorological Society, 1997, 78(2): 197–208, https://doi.org/10.1175/1520-0477(1997)078%3C0197:EAGMEB%3E2.0.CO;2
    [37] LI J D, GEEN R, MAO J Y, et al. Mechanical and thermal forcings of Asian large-scale orography on spring cloud amount and atmospheric radiation budget over East Asia [J]. Journal of Climate, 2023, 36(15): 5215–5232, https://doi.org/10.1175/JCLI-D-22-0797.1
    [38] LIU Y Y, CHEN L J. Features and possible causes for the spring climate anomalies in 2019 [J]. Meteorological Monthly, 2019, 45(10): 1483–1493, in Chinese with English abstract
    [39] DU J, FANG J, XU W, et al. Analysis of dry/wet conditions using the standardized precipitation index and its potential usefulness for drought/flood monitoring in Hunan Province, China [J]. Stochastic Environmental Research and Risk Assessment, 2013, 27(2): 377–387, https://doi.org/10.1007/s00477-012-0589-6
    [40] VICENTE-SERRANO S M, BEGUERÍA S, LÝPEZ-MORENO J I. A multiscalar drought index sensitive to global warming: the standardized precipitation evapotran-spiration index [J]. Journal of Climate, 2010, 23(7): 1696-–1718, https://doi.org/10.1175/2009JCLI2909.1
    [41] ALLEN R G, PEREIRA L S, RAES D, et al. Crop evapotranspiration -Guidelines for computing crop requirements-FAO Irrigation and drainage paper 56 [J]. Food and Agriculture Organization, Rome, 1998, 300(9): D05109
    [42] VALIANTZAS J D. Simplified versions for the Penman evaporation equation using routine weather data [J]. Journal of Hydrology, 2006, 331(3): 690–702, https://doi.org/10.1016/j.jhydrol.2006.06.012
    [43] MCCOLL K A. Practical and theoretical benefits of an alternative to the Penman‐Monteith evapotranspiration equation [J]. Water Resources Research, 2020, 56(6): e2020WR027106, https://doi.org/10.1029/2020WR027106
    [44] HOMDEE T, PONGPUT K, KANAE S. A comparative performance analysis of three standardized climatic drought indices in the Chi River basin, Thailand [J]. Agriculture and Natural Resources, 2016, 50(3): 211–219, https://doi.org/10.1016/j.anres.2016.02.002
    [45] WANG Q F, ZENG J Y, QI J Y, et al. A multi-scale daily SPEI dataset for drought characterization at observation stations over mainland China from 1961 to 2018 [J]. Earth System Science Data, 2021, 13(2): 331–341, https://doi.org/10.5194/essd-13-331-2021
    [46] CHEN D, GAO G, XU C Y, et al. Comparison of the Thornthwaite method and pan data with the standard Penman-Monteith estimates of reference evapotranspiration in China [J]. Climate Research, 2005, 28(2): 123–132, https://doi.org/10.3354/cr028123
    [47] HERSBACH H, BELL B, BERRISFORD P, et al. The ERA5 global reanalysis [J]. Quarterly Journal of the Royal Meteorological Society, 2020, 146(730): 1999–2049, https://doi.org/10.1002/qj.3803
    [48] AHLGREN P, JARNEVING B, ROUSSEAU R. Requirements for a cocitation similarity measure, with special reference to Pearson's correlation coefficient [J]. Journal of the American Society for Information Science and Technology, 2003, 54(6): 550–560, https://doi.org/10.1002/asi.10242
    [49] PEARSON K. LⅢ On lines and planes of closest fit to systems of points in space [J]. Philosophical Magazine, 1901, 2(11): 559‒572, https://doi.org/10.1080/14786440109462720
    [50] NI Y, HSU P C. Inter-annual variability of global monsoon precipitation in present-day and future warming scenarios based on 33 Coupled Model Intercomparison Project Phase 5 models [J]. International Journal of Climatology, 2018, 38(13): 4875–4890, https://doi.org/10.1002/joc.5704
    [51] HE C, LI T. Does global warming amplify interannual climate variability [J]. Climate Dynamics, 2019, 52(5–6): 2667–2684, https://doi.org/10.1007/s00382-018-4286-0
    [52] HOSKINS B, PEDDER M, JONES D W. The omega equation and potential vorticity [J]. Quarterly Journal of the Royal Meteorological Society, 2003, 129(595): 3277–3303, https://doi.org/10.1256/qj.02.135
    [53] WU G X, MA T T, LIU Y M, et al. PV-Q perspective of cyclogenesis and vertical velocity development downstream of the Tibetan Plateau [J]. Journal of Geophysical Research: Atmospheres, 2020, 125(16): e2019JD030912, https://doi.org/10.1029/2019JD030912
    [54] HE C. Future drying subtropical East Asia in Winter: mechanism and observational constraint [J]. Journal of Climate, 2023, 36(9): 2985–2998, https://doi.org/10.1175/JCLI-D-22-0347.1
    [55] KIM E J, HONG S Y. Impact of air-sea interaction on East Asian summer monsoon climate in WRF [J]. Journal of Geophysical Research Atmospheres, 2010, 115: D19118, https://doi.org/10.1029/2009JD013253
    [56] CHEN W, LEE J Y, LU R Y, et al. Intensified impact of tropical Atlantic SST on the western North Pacific summer climate under a weakened Atlantic thermohaline circulation [J]. Climate Dynamics, 2015, 45(7): 2033–2046, https://doi.org/10.1007/s00382-014-2454-4

Get Citation+

ZHAO Qing-hong, YANG Song, TIAN Hong-ying, et al. Leading Pattern of Spring Drought Variability over East Asia and Associated Drivers [J]. Journal of Tropical Meteorology, 2024, 30(1): 1-10, https://doi.org/10.3724/j.1006-8775.2024.001
ZHAO Qing-hong, YANG Song, TIAN Hong-ying, et al. Leading Pattern of Spring Drought Variability over East Asia and Associated Drivers [J]. Journal of Tropical Meteorology, 2024, 30(1): 1-10, https://doi.org/10.3724/j.1006-8775.2024.001
Export:  

Share Article

Manuscript History

Manuscript received: 07 October 2023
Manuscript revised: 15 November 2023
Manuscript accepted: 15 February 2024
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

Leading Pattern of Spring Drought Variability over East Asia and Associated Drivers

doi: 10.3724/j.1006-8775.2024.001
Funding:

National Natural Science Foundation of China 42230603

National Natural Science Foundation of China 42275020

Guangdong Major Project of Basic and Applied Basic Research 2020B0301030004

Guangdong Province Key Laboratory for Climate Change and Natural Disaster Studies 2020B1212060025

Innovation Group Project of Southern Marine Science and Engineering Guangdong Laboratory (Zhuhai) 311021001

Open Fund of State Key Laboratory of Satellite Ocean Environment Dynamics, Second Institute of Oceanography, MNR QNHX2310

Abstract: Drought events have become more frequent and intense over East Asia in recent decades, leading to huge socioeconomic impacts. Although the droughts have been studied extensively by cases or for individual regions, their leading variability and associated causes remain unclear. Based on the Standardized Precipitation Evapotranspiration Index (SPEI) and ERA5 reanalysis product from 1979 to 2020, this study evealuates the severity of spring droughts in East Asia and investigates their variations and associated drivers. The results indicate that North China and Mongolia have experienced remarkable trends toward dryness during spring in recent decades, while southwestern China has witnessed an opposite trend toward wetness. The first Empirical Orthogonal Function mode of SPEI variability reveals a similar seesawing pattern, with more severe dryness in northwestern China, Mongolia, North China, South Korea, and Japan but increased wetness in Southwestern China and southeast Asia. Further investigation reveals that the anomalously dry (wet) surface in North (Southwestern) China is significantly associated with anomalously high (low) temperature, less (more) precipitation, and reduced (increased) soil moisture during the previous winter and early spring, regulated by an anomalous anticyclone (cyclone) and thus reduced (increased) water vapor convergence. The spring dry-wet pattern in East Asia is also linked to cold sea surface temperature anomalies in the central-eastern Pacific. The findings of this study have important implications for improving the prediction of spring drought events in East Asia.

ZHAO Qing-hong, YANG Song, TIAN Hong-ying, et al. Leading Pattern of Spring Drought Variability over East Asia and Associated Drivers [J]. Journal of Tropical Meteorology, 2024, 30(1): 1-10, https://doi.org/10.3724/j.1006-8775.2024.001
Citation: ZHAO Qing-hong, YANG Song, TIAN Hong-ying, et al. Leading Pattern of Spring Drought Variability over East Asia and Associated Drivers [J]. Journal of Tropical Meteorology, 2024, 30(1): 1-10, https://doi.org/10.3724/j.1006-8775.2024.001
  • Drought events, characterized by prolonged periods of abnormally low precipitation (Chen et al.[1]; Li et al.[2]), can exert significant impacts on the hydrological cycles, ecosystems, and human society (Huang et al.[3]; Chen et al.[4]; Kostopoulou and Giannakopoulos[5]). The occurrence of severe droughts not only affects crop growth and agricultural production in the spring (Han et al.[6]; Santini et al.[7]; Higgins et al.[8]), but also leads to an increased frequency of natural disasters such as heatwaves and wildfires in summer due to dry surface conditions (Purich et al.[9]; Wang et al.[10]; Heino et al.[11]). In recent decades, drought events have become more frequent across the world, resulting in substantial socioeconomic impacts (Suska and Piotr[12]; Li et al.[13]; Dong et al.[14]). Notably, East Asia is a region particularly susceptible to the drought disasters (Zhou and Wang[15]; Deng et al.[16]). For instance, Southwestern China experienced unprecedented droughts during the spring of 2010, resulting in water scarcity, crop failures, and significant economic losses (Zhang et al.[17]; Liu et al.[18]). Between July and August 2022, the Yangtze River basin of China faced its most severe drought on record, causing water levels in China's largest freshwater lake, Poyang Lake, to plummet from 19 meters in June to 9 meters by the end of August (Mallapaty[19]). Moreover, the frequency and severity of extreme dry events are projected to increase in a warming climate (Qiao et al.[20]; Tang et al.[21]; Ullah et al.[22]). Hence, it is important to enhance our understanding of the changes and variations in drought events over East Asia, as well as their underlying physical causes.

    The drought severity in East Asia is significantly influenced by the Asian monsoon system (Yu et al.[23]; Cen et al.[24]; Huang et al.[25]). Previous studies have indicated that the East Asian summer monsoon circulation has weakened in recent decades (Wang[26]; Wang et al.[27]; Zhou and Yu[28]), causing reduced precipitation in North China and consequently exacerbating the dry conditions there (Li et al.[29]; Zhou et al.[30]; Zhang et al.[31]). In contrast, the changes in the South Asian summer monsoon can affect the transport of water vapor from the tropical Indian Ocean through the Bay of Bengal to Southwestern China, affecting the dryness in the region (Simmonds et al.[32]; Zhou and Yu[28]). Moreover, the global SST anomalies can significantly regulate the interannual to interdecadal variations of the drought events in East Asia. Particularly, the El Niño-Southern Oscillation (ENSO) plays a crucial role in driving the interannual variations of extreme dry conditions in China (Cao et al.[33]; Wang and Li[34]). The El Niño-like SST can lead to the intensification and westward shift of the Northwestern Pacific subtropical high, resulting in reduced precipitation and increased temperatures in South and Southwestern China, which in turn favor the occurrence of drought events (Wang et al.[35]). In addition, Asian large-scale orography forcing, particularly the Tibetan Plateau (TP), significantly increases ascent, low-middle cloud formation, and resultant strong spring cloud radiative cooling over Southeastern China (SEC) and downstream ocean, thereby affecting climate change (Kiehl and Trenberth[36]; Li et al.[37]). Studies suggest that the western Pacific subtropical high anomaly was stronger, westward and southward in the spring of 2019, which directly led to significantly more precipitation in Northeastern China, eastern Northwestern China, and South China, while the Huang-Huai, Jiang-Huai and most Yunnan provinces had less precipitation (Liu and Chen[38]).

    Although previous studies have extensively explored the dry and wet extremes in East Asia (Du et al.[39]; Huang et al.[3]; Qiao et al.[20]), most are focused on the trends, on individual cases, or for specific regions. The leading mode of drought variability across East Asia during the boreal spring and its underlying causes remain unclear. Since spring marks the period of crop development in East Asia, alterations in springtime drought conditions can significantly affect crop growth, posing a potential threat to regional food prices and security. Therefore, this study aims to assess the severity of drought in East Asia and investigate its variations and physical causes during the boreal spring. Section 2 presents the data and methods used in this research. Sections 3 and 4 present the research findings, followed by a summary and discussion in Section 5.

  • The Standardized Precipitation Evapotranspiration Index (SPEI) is a multiscalar drought index derived from climatic data. It serves as a tool for identifying the onset, duration, and severity of drought conditions in comparison to typical circumstances (Vicente-Serrano et al.[40]). Mathematically, SPEI is simply computed by total precipitation minus potential evapotranspiration. However, the estimation of potential evapotranspiration is difficult in regions without sufficient data, which follows the Penman-Monteith method (Allen et al.[41]):

    $$ \mathrm{ET}_0=\frac{0.408 \Delta\left(R_n-G\right)+\gamma \frac{900}{T+273} u_2\left(e_s-e_a\right)}{\Delta+\gamma\left(1+0.34 u_2\right)} $$ (1)

    In Formula 1, ET0 is the potential evapotranspiration (mm d–1), and Rn is the net radiation (MJ m–2 d–1). G is soil heat flux (MJ m–2 d–1); T is the daily average temperature (℃); u2 is the wind speed at 2 m (m s–1), obtained by multiplying the wind speed at 10 m by 0.75. es is saturated vapor pressure (kPa); ea is the actual water vapor pressure (kPa); Δ is the slope of saturated water vapor pressure-temperature curve (kPa ℃–1); γ is a hygroscopic constant (kPa ℃–1) (Valiantzas[42]; McColl[43]). SPEI offers a comprehensive assessment of drought by taking into account both evapotranspiration and precipitation, enabling the detection of drought across various temporal scales (Vicente-Serrano et al.[40]; Homdee et al.[44]; Wang et al.[45]). In the context of studies in China, the Penman-Monteith method has been found to be particularly suitable (Chen et al.[46]).

  • ERA5 is the state-of-the-art reanalysis dataset, which is generated through the 4D-Var data assimilation of the ECMWF Integrated Forecast System (IFS) and the CY41R2 model (Hersbach et al.[47]). In this study, the variables that we analyze in this study include 2-m temperature, precipitation, soil moisture (at a depth of 1 m underground), three-dimensional winds, geopotential height, surface pressure, sea surface temperature, specific humidity, total cloud cover, and solar downward short-wave radiation. The horizontal resolution of these variables is 1°×1° (longitude × latitude) for the period from 1979 to 2020. The study area of this study, i.e., East Asia, spans from 15°N to 55°N and from 70°E to 150°E.

  • In this study, various statistical methods have been employed to investigate the characteristics and the associated drivers of spring drought severity across East Asia. These methods include the Empirical Orthogonal Function (EOF) analysis, correlation/regression analysis, composite analysis, trend analysis, and lagged correlation/regression analysis. The correlation coefficient (r) is calculated as follows:

    $$ r=\frac{\sum_{i=1}^n\left(x_i-\bar{x}\right)\left(y_i-\bar{y}\right)}{\sqrt{\sum_{i=1}^n\left(x_i-\bar{x}\right)^2 \sum_{i=1}^n\left(y_i-\bar{y}\right)^2}} $$ (2)

    where x1, x2, ..., xn and y1, y2, ..., yn is a random variable for two groups of data (Ahlgren et al.[48]). The range of correlation coefficient r is between [–1, 1]. The greater the absolute value of r, the greater the correlation between the two sets of data.

  • Figure 1 shows the climatology, trend, and standardized deviation of SPEI in the spring for the period 1979–2020. The climatology of SPEI shows that Northwestern China, North China, and Mongolia typically featured an arid climate (Fig. 1a). The trend in SPEI reveals that Northwestern China, North China, and Mongolia experienced a drying trend over the past several decades, while Southwestern China displayed a wetting trend (Fig. 1b). The standardized deviation of SPEI is relatively large in Northwestern China, North China, and Southwestern China, implying that there exists a larger variability of drought in these regions (Fig. 1c).

    Figure 1.  (a) Climatology, (b) linear trend, and (c) standardized deviation of the spring SPEI over East Asia.

    Figure 2 further presents the first EOF mode (EOF1) of SPEI variability in East Asia during the boreal spring. EOF analysis is also known as principal component analysis (PCA) (Pearson[49]). The first EOF mode (EOF1) has the most obvious evolution of spatial distribution characteristics and coefficients over time and has the largest interpretation variance. Only EOF1 is analyzed here. The EOF1 shows a north-south dipolar pattern, with opposite signs of SPEI anomalies between northeastern and southwestern parts of East Asia (Fig. 2a). Moreover, the primary component of EOF1 (PC1) shows an overall upward trend, generally changing from negative phases to positive phases (Fig. 2b). Combined with the spatial characteristics of EOF1, it indicates that Northwestern China, North China, central China, and Mongolia have experienced a drying trend, while Southwestern China has experienced a wetting trend.

    Figure 2.  (a) Spatial pattern and (b) primary component of EOF1 of variability in the spring SPEI over East Asia. In (b), the black curve and dashed line indicate the original PC1 and its regression trends, respectively, while the bars represent the PC1 with the trends removed.

    Figure 3 shows the SPEI composite results of extremely dry and wet years based on a 1.5 standardized deviation of the de-trended PC1. The years with standard deviations higher than 1.5 are characterized by dry surface conditions in Northwestern China, North China, and Mongolia and wet surface conditions in Southwestern China (Fig. 3a). The opposite results were obtained for the years with standard deviations below 1.5 (Fig. 3b). The differences in SPEI between the positive and negative phases of PC1 further highlight the dry-wet pattern, where the absolute value of the SPEI differences can be higher than 4.5 in North China and Southwestern China (Fig. 3c).

    Figure 3.  Composite patterns of spring SPEI are based on the years when the standardized deviations of PC1 are (a) more and (b) less than 1.5. (c) Difference between (a) and (b).

  • Figure 4 displays the lagged correlations between PC1 and the SPEI. The signal of the East Asian dry-wet patterns appeared four months before spring. In periods with two- and zero-month advance, the correlation coefficients between PC1 and SPEI become more significant, with the main dry areas occurring in Northwestern China, North China, and Mongolia, while wet areas in Southwestern China.

    Figure 4.  Lag-lead correlations between SPEIs in different lags and the detrend PC1 during 1979–2020.

    The SPEI variations are closely related to the temperature and precipitation. As illustrated by Fig. 5, the anomalous dryness in Northwestern China, parts of North China, and Southern Mongolia corresponds to increased air temperature (left panels) and decreased precipitation and soil moisture (middle and right panels, respectively) four and two months preceding the spring drought (left panels). By comparison, the anomalous wetness in Southwestern China is correlated with abnormal temperature and precipitation patterns, opposite to those in the dry areas. These significant lagged correlation patterns also imply that the dry-wet variations of East Asian drought, as revealed by the EOF1, possess a potential for subseasonal and seasonal predictions.

    Figure 5.  Lag-lead correlations of 2-m temperature (left), precipitation (middle), and soil moisture (right) in different leads with the detrend PC1.

    Figure 6 shows the spatial distributions of regressions of geopotential height and horizontal winds onto the PC1 over East Asia. The results show that preceding the spring dry and wet variations, an abnormal high pressure accompanied by anticyclonic circulation occurred over Northwestern China and Mongolia, and an abnormal low pressure accompanied by cyclonic circulation existed over Japan. North China is located between the anomalous high and low pressure systems and is affected by northerly wind anomalies.

    Figure 6.  Lagged regressions of geopotential height (units: dagpm) and horizontal winds (units: m s–1) in different leads at 850 hPa (left), 500 hPa (middle), and 300 hPa (right) onto the detrend PC1. Areas with significant anomalies exceeding the 95% confidence level are stippled.

    Figure 7 further presents the regression patterns of water vapor flux, total cloud cover, and solar shortwave radiation onto PC1. The results show that the water vapor flux anomalies are closely linked to the wind anomalies. The drying areas such as Northwestern China, North China, and Mongolia are primarily featured by divergent anomalies of water vapor fluxes. The variability of precipitation is directly controlled by local vertical velocity anomaly, and anomalous ascent (descent) leads to excessive (deficient) precipitation (Ni and Hsu[50]; He and Li[51]). According to isentropic gliding mechanism, the climatological isentropic surfaces tilt northward with altitude in the extratropics, and anomalous northerly (southerly) wind generates descent (ascent) anomaly by gliding along the sloping isentropic surfaces (Hoskins et al.[52]; Wu et al.[53]; He[54]), associated with anomalous divergence (convergence) in lower troposphere. Therefore, the dry areas are accompanied by water vapor divergence, reduced total cloud cover, and increased short-wave solar radiation due to the anomalous anticyclone and stronger descending motions. By comparison, the anomalous wetness in Southwestern China has obtained increased water vapor transport from the north Pacific, increased total cloud cover, and decreased solar radiation. As a result, the increased surface air temperature and reduced precipitation favor anomalous dry conditions in Northwestern China and North China and vice versa in Southwestern China.

    Figure 7.  Regressions of the vertical integrals of (a) water vapor flux (vectors, units: kg m–1 s–1) and divergence (units: kg m–2 s–1), (b) total cloud cover, and (c) solar download radiation in spring onto the detrend PC1. Areas with significant anomalies exceeding the 95% confidence level are stippled.

  • The tropical SST anomalies can exert significant impacts on weather and climate extremes in East Asia. Fig. 8a displays the correlation coefficients between PC1 and the SST, which shows that the spatial patterns of dry North China and wet Southwestern China are correlated with the La Niña - like SST anomalies, coinciding with warm SST anomalies in the western Pacific and cold SST anomalies in the central and eastern Pacific and the Indian Ocean. Consequently, the SST gradients between the western Pacific and the central and eastern Pacific increase. Fig. 8b shows the regression patterns of PC1 with geopotential height and horizontal winds, which show anomalous high and low pressures over East Asia, which resemble a Rossby wave-train linked to the SST anomalies. When the western Pacific Sea temperature rises, convective activities are motivated, abnormal high pressure will be generated over the northwest of Mongolia, Alaska, and the western coastal areas of the United States, while obvious abnormal low pressure will appear over Japan, forming a "+–+" teleconnection wave train propagating from west to east.

    Figure 8.  (a) Regression between SST in spring and the detrend PC1. (b) Regression of 300-hPa geopotential height and horizontal winds onto the detrend PC1.

    Previous studies have indicated that SST changes in the Northern Indian Ocean and the tropical Northern Atlantic affect circulation systems in the Northwestern Pacific region through air-sea interactions, which has a significant impact on the weather and climate in East Asia (Kim and Hong[55]; Chen et al.[56]). The western Pacific (10°N–20°N, 120°E–180°), the middle Eastern Pacific (–5°S–15°N, 120°W–160°W), and the Northern Indian Ocean (–10°S–10°N, 70°E–120°E) are thus selected as three key ocean areas to define a new SST index:

    $$ S=2 \times \mathrm{SST}_1-\mathrm{SST}_2-\mathrm{SST}_3 $$ (3)

    where subscripts 1, 2, and 3, respectively, represent the mean SST over the three selected sea areas. The correlation coefficient r between S and PC1 is about 0.514, which has passed the significance test (Fig. 9a), indicating that the tropical zonal SST gradient in the Indo-Pacific region is significantly correlated with the spring dry and wet variations in East Asia. When the Indo-Pacific SST gradient increases, dry (wet) surface conditions tend to occur in Northwestern China, North China, and Mongolia (Southwestern China). In addition, the results of Figs. 9c and 9d are consistent with those of Fig. 7a and Fig. 4e, indicating that the tropical Pacific SST anomalies could significantly influence the spring dry and wet variations over East Asia by triggering atmospheric circulations.

    Figure 9.  (a) SST index averaged over the area in the box shown in Fig. 8a (bars) and the detrend PC1 (curve). (b), (c), and (d) respectively represent the regressions of 300-hPa geopotential height and horizontal winds, the water vapor flux and divergence, and the SPEI onto the SST index.

  • Based on the global SPEI data calculated by the Penman-Monteith method and the ERA5 reanalysis product from 1979 to 2020, this study investigates the characteristics and possible drivers of the spring dry and wet variations in East Asia. The main conclusions are as follows: (1) Over the past several decades, the leading variability in spring SPEI over East Asia has been characterized by a dipole pattern, with increased dryness (wetness) in Northwestern China, North China, and Mongolia (Southwestern China). (2) The leading mode of SPEI variability in East Asia is significantly correlated with the anomalous anticyclone (cyclone) over North China and Mongolia (Southwestern China), leading to increased (decreased) temperature and decreased (increased) precipitation and soil moisture in North China and Mongolia (Southwestern China). (3) The tropical Pacific SST anomalies could influence the leading mode of SPEI variability in East Asia during boreal spring by triggering anomalous atmospheric circulation.

    It should be noted that our research shows that the variability of spring dryness in East Asia signals 2–4 months in advance, which provides some help in predicting sub-seasonal and seasonal changes in the regional spring drought. Although our study suggests that the anomalous SSTs in the tropical Pacific may influence the spring drought in East Asia, these findings are primarily based on statistical analyses. Therefore, validation through numerical modeling experiments is critical for further substantiating the possible drivers. Furthermore, in addition to the tropical Pacific Ocean, the Northern Atlantic SST, and atmospheric variability, such as the Atlantic Multidecadal Oscillation and the North Atlantic Oscillation, could also influence the spring drought in East Asia. However, how these factors interact synergistically with the Pacific SSTs remains unclear and requires further in-depth analyses in the future.

Reference (56)

Catalog

    /

    DownLoad:  Full-Size Img  PowerPoint
    Return
    Return