Advanced Search
Article Contents

High-Resolution Modeling Study of the Kuroshio Path Variations South of Japan

Fund Project:

doi: 10.1007/s00376-014-3230-4

  • A high-resolution ocean general circulation model (OGCM) is used to investigate the Kuroshio path variations south of Japan. The model reproduces many important features of the Kuroshio system including its interannual bimodal variability south of Japan. A decreasing trend of the spatial averaged relative vorticity is detected when the Kuroshio takes the non-large meander (NLM) path, and during the transition period from the NLM to the large meander (LM), a sudden release of velocity shear corresponds well to the weakening of the Shikoku recirculation gyre (SRG), which plays a key role in modulating the Kuroshio path variations. Analysis of eddy energetics indicates that baroclinic instability is mainly responsible for the formation of the LM. In addition, further analysis shows that the strength of the SRG could be largely influenced by the baroclinic Rossby wave adjustment process, forced by the wind stress curl anomalies in the North Pacific basin, based on the model investigation. It is suggested that the cyclonic disturbances might account for the weakening of the SRG, and act as a remote trigger for the baroclinic instability of the Kuroshio south of Japan.
    摘要: A high-resolution ocean general circulation model (OGCM) is used to investigate the Kuroshio path variations south of Japan. The model reproduces many important features of the Kuroshio system including its interannual bimodal variability south of Japan. A decreasing trend of the spatial averaged relative vorticity is detected when the Kuroshio takes the non-large meander (NLM) path, and during the transition period from the NLM to the large meander (LM), a sudden release of velocity shear corresponds well to the weakening of the Shikoku recirculation gyre (SRG), which plays a key role in modulating the Kuroshio path variations. Analysis of eddy energetics indicates that baroclinic instability is mainly responsible for the formation of the LM. In addition, further analysis shows that the strength of the SRG could be largely influenced by the baroclinic Rossby wave adjustment process, forced by the wind stress curl anomalies in the North Pacific basin, based on the model investigation. It is suggested that the cyclonic disturbances might account for the weakening of the SRG, and act as a remote trigger for the baroclinic instability of the Kuroshio south of Japan.
  • 加载中
  • Akitomo, K.,M. Ooi,T. Awaji, and K. Kutsuwada, 1996:Interannual variability of the Kuroshio transport in response to the wind stress field over the North Pacific: Its relation to the path variation south of Japan.J. Geophys. Res.,101,14 057-14 071.
    Andres, M.,J. H. Park,M. Wimbush,X. H. Zhu,K. I. Chang, and H. Ichikawa, 2008:Study of the Kuroshio/Ryukyu Current System based on satellite-altimeter and in situ measurements.J. Oceanogr.,64,937-950.
    Bryan, K., 1969:A numerical method for the study of the circulation of the world ocean.J. Comput. Phys.,4,347-376.
    Cessi, P.,G. Ierley, and W. Young, 1987:A model of the inertial recirculation driven by potential vorticity anomalies.J. Phys. Oceanogr.,17,1640-1652.
    Chao, S. Y., and J. P. McCreary, 1982:A numerical study of the Kuroshio south of Japan.J. Phys. Oceanogr.,12,679-693.
    Cox, M. D., 1984:A primitive equation, three-dimensional model of the ocean.GFDL Ocean Group Tech. Rep. No. 1,GFDL,Princeton, NJ, USA, 143 pp.
    Douglass, E. M.,S. R. Jayne,F. O. Bryan,S. Peacock, and M. Maltrud, 2012:Kuroshio pathways in a climatologically forced model.J. Oceanogr.,68,625-639.
    Endoh, T., and T. Hibiya, 2001:Numerical simulation of the transient response of the Kuroshio leading to the large meander formation south of Japan.J. Geophys. Res.,106,26 833-26 850.
    Endoh, T., and T. Hibiya, 2009:Interaction between the trigger meander of the Kuroshio and the abyssal anticyclone over Koshu seamount as seen in the reanalysis data.Geophys. Res. Lett.,36,L18604, doi: 10.1029/2009GL039389.
    Griffies, S. M., and Coauthors, 2009:Coordinated ocean-ice reference experiments (COREs).Ocean Modelling,26,1-46.
    Hurlburt, H. E.,A. J. Wallcraft, W. J. Schmitz Jr.,P. J. Hogan, and E. J. Metzger, 1996:Dynamics of the Kuroshio/Oyashio current system using eddy-resolving models of the North Pacific Ocean.J. Geophys. Res.,101,941-976.
    Ierley, G. R., and W. R. Young, 1988:Inertial recirculation in a β-plane corner.J. Phys. Oceanogr.,18,683-689.
    Jacob, R. L., 1997:Low frequency variability in a simulated atmosphere ocean system.Ph. D dissertation,University of Wisconsin-Madison,159 pp.
    Jensen, T. G., 1996:Artificial retardation of barotropic waves in layered ocean models.Mon. Wea. Rev.,124,1272-1284.
    Kawabe, M., 1980:Sea level variations around the Nansei Islands and the large meander in the Kuroshio south of central Japan.J. Oceanogr. Soc. Japan,36,227-235.
    Kawabe, M., 1985:Sea level variations at the Izu Islands and typical stable paths of the Kuroshio.J. Oceanogr. Soc. Japan,41,307-326.
    Kawabe, M., 1986:Transition processes between the three typical paths of the Kuroshio.J. Oceanogr. Soc. Japan.,42,174-191.
    Kawabe, M., 1995:Variations of current path, velocity, and volume transport of the Kuroshio in relation with the large meander.J. Phys. Oceanogr.,25,3103-3117.
    Killworth, P. D.,D. Stainforth,D. J. Webb, and S. M. Paterson, 1991:The development of a free-surface Bryan-Cox-Semtner ocean model.J. Phys. Oceanogr.,21,1333-1348.
    Large, W. G., and S. G. Yeager, 2004: Diurnal to decadal global forcing for ocean and Sea-Ice models: the data sets and flux Climatologies. NCAR/TN-460+STR, 1-105.
    Large, W. G., and S. G. Yeager, 2009:The global climatology of an interannually varying air-sea flux data set.Climate Dyn.,33,341-364.
    Liu, Z.Y., 1997:The influence of stratification on the inertial recirculation.J. Phys. Oceanogr.,27,926-940.
    Maltrud, M. E., and J. L. McClean, 2005:An eddy resolving global 1/10° ocean simulation.Ocean Modelling,8,31-54.
    Marchesiello, P.,J. C. McWilliams, and A. F. Shchepetkin, 2001:Open boundary conditions for long-term integration of regional oceanic models.Ocean Modelling,3,1-20.
    Mitsudera, H.,T. Waseda,Y. Yoshikawa, and B. Taguchi, 2001:Anticyclonic eddies and Kuroshio meander formation.Geophys. Res. Lett.,28,2025-2028.
    Miyazawa, Y.,X. Guo, and T. Yamagata, 2004:Roles of meso-scale eddies in the Kuroshio paths.J. Phys. Oceanogr.,34,2203-2222.
    Miyazawa, Y.,T. Kagimoto,X. Guo, and H. Sakuma, 2008:The Kuroshio large meander formation in 2004 analyzed by an eddy-resolving ocean forecast system.J. Geophys. Res.,113,C10015, doi: 10.1029/2007JC004226.
    Nagano, A.,K. Ichikawa,H. Ichikawa,H. Tomita,H. Tokinaga, and M. Konda, 2010:Stable volume and heat transports of the North Pacific subtropical gyre revealed by identifying the Kuroshio in synoptic hydrography south of Japan.J. Geophys. Res.,115,C09002, doi: 10.1029/2009JC005747.
    Pacanowski, R., 1995:MOM2 documentation, user's guide and reference manual.GFDL Ocean. Tech. Rep. No. 3,Geophys. Fluid Dyn. Lab.,Princeton, N. J., 329 pp.
    Qiu, B., 2003:Kuroshio Extension variability and forcing of the Pacific decadal oscillations: Responses and potential feedback.J. Phys. Oceanogr.,33,2465-2482.
    Qiu, B., and T. M. Joyce, 1992:Interannual variability in the mid- and low-latitude western North Pacific.J. Phys. Oceanogr.,22,1062-1079.
    Qiu, B., and W. F. Miao, 2000:Kuroshio path variations south of Japan: Bimodality as a self-sustained internal oscillation.J. Phys. Oceanogr.,30,2124-2137.
    Qiu, B., and S. M. Chen, 2005:Variability of the Kuroshio Extension jet, recirculation gyre and mesoscale eddies on decadal timescales.J. Phys. Oceanogr.,35,2090-2103.
    Semtner, A. J., 1974: A general circulation model for the World Ocean. UCLA Dept. of Meteor. Tech. Rep. No. 8, 99 pp.
    Taft, B. A., 1972:Characteristics of the flow of the Kuroshio south of Japan.Kuroshio-Its Physical Aspects,H. Stommel and K. Yoshida,Eds., University of Tokyo Press, Tokyo, 165-214.
    Tobis, M., 1996:Effect of slowed barotropic dynamics in parallel ocean climate models.Ph. D. dissertation,University of Wisconsin-Madison,198 pp.
    Tobis, M.,C. Schafer,I. Foster,R. Jacob, and J. Anderson, 1997:FOAM: Expanding the horizons of climate modeling.Proc. Super Computing 1997,San Diego,California, ACM/IEEE, 27 pp.
    Tsujino, H.,N. Usui, and H. Nakano, 2006:Dynamics of Kuroshio path variations in a high-resolution general circulation model.J. Geophys. Res.,111,C11001, doi: 10.1029/2005JC003118.
    Tsujino, H.,S. Nishikawa,K. Sakamoto,N. Usui,H. Nakano, and G. Yamanaka, 2013:Effects of large-scale wind on the Kuroshio path south of Japan in a 60-year historical OGCM simulation.Climate Dyn.,41,2287-2318.
    Usui, N.,H. Tsujino,H. Nakano, and Y. Fujii, 2008:Formation process of the Kuroshio large meander in 2004.J. Geophys. Res.,113,C08047, doi: 10.1029/2007JC004675.
    Usui, N.,H. Tsujino,H. Nakano,Y. Fujii, and M. Kamachi, 2011:Decay mechanism of the 2004/05 Kuroshio large meander.J. Geophys. Res.,116,C10010, doi: 10.1029/2011JC007009.
    Waseda, T.,H. Mitsudera,B. Taguchi, and Y. Yoshikawa, 2003:On the eddy-Kuroshio interaction: Meander formation process.J. Geophys. Res.,108(C7),3220, doi: 10.1029/2002JC001583.
    White, W. B., and J. P. McCreary, 1976:On the formation of the Kuroshio meander and its relationship to the large-scale ocean circulation.Deep Sea Res.,23,33-47.
    Yoon, J.-H., and I. Yasuda, 1987:Dynamics of the Kuroshio large meander: Two-layer model.J. Phys. Oceanogr.,17,66-81.
  • [1] 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
    [2] Hanying LI, Peng HU, Qiong ZHANG, Ashish SINHA, Hai CHENG, 2021: Understanding Interannual Variations of the Local Rainy Season over the Southwest Indian Ocean, ADVANCES IN ATMOSPHERIC SCIENCES, 38, 1852-1862.  doi: 10.1007/s00376-021-1065-3
    [3] Ran LIU, Changlin CHEN, Guihua WANG, 2016: Change of Tropical Cyclone Heat Potential in Response to Global Warming, ADVANCES IN ATMOSPHERIC SCIENCES, 33, 504-510.  doi: 10.1007/s00376-015-5112-9
    [4] Xiao DONG, Jiangbo JIN, Hailong LIU, He ZHANG, Minghua ZHANG, Pengfei LIN, Qingcun ZENG, Guangqing ZHOU, Yongqiang YU, Mirong SONG, Zhaohui LIN, Ruxu LIAN, Xin GAO, Juanxiong HE, Dongling ZHANG, Kangjun CHEN, 2021: CAS-ESM2.0 Model Datasets for the CMIP6 Ocean Model Intercomparison Project Phase 1 (OMIP1), ADVANCES IN ATMOSPHERIC SCIENCES, 38, 307-316.  doi: 10.1007/s00376-020-0150-3
    [5] ZHU Jieshun, SUN Zhaobo, ZHOU Guangqing, 2007: A Note on the Role of Meridional Wind Stress Anomalies and Heat Flux in ENSO Simulations, ADVANCES IN ATMOSPHERIC SCIENCES, 24, 729-738.  doi: 10.1007/s00376-007-0729-y
    [6] LI Chun, SUN Jilin, 2015: Role of the Subtropical Westerly Jet Waveguide in a Southern China Heavy Rainstorm in December 2013, ADVANCES IN ATMOSPHERIC SCIENCES, 32, 601-612.  doi: 10.1007/s00376-014-4099-y
    [7] LIU Yudi, WANG Bin, JI Zhongzhen, 2003: Research on Atmospheric Motion in Horizontal Discrete Grids, ADVANCES IN ATMOSPHERIC SCIENCES, 20, 139-148.  doi: 10.1007/BF03342058
    [8] Tong XIE, Liguang WU, Yecheng FENG, Jinghua YU, 2024: Alignment of Track Oscillations during Tropical Cyclone Rapid Intensification, ADVANCES IN ATMOSPHERIC SCIENCES, 41, 655-670.  doi: 10.1007/s00376-023-3073-y
    [9] Huang Ronghui, Zang Xiaoyun, Zhang Renhe, Chen Jilong, 1998: The Westerly Anomalies over the Tropical Pacific and Their Dynamical Effect on the ENSO Cycles during 1980-1994, ADVANCES IN ATMOSPHERIC SCIENCES, 15, 135-151.  doi: 10.1007/s00376-998-0035-3
    [10] YU Yongqiang, ZHANG Xuehong, GUO Yufu, 2004: Global Coupled Ocean-Atmosphere General Circulation Models in LASG/IAP, ADVANCES IN ATMOSPHERIC SCIENCES, 21, 444-455.  doi: 10.1007/BF02915571
    [11] WANG Yafei, Fujiyaoshi YASUSHI, Kato KURANOSHIN, 2003: A Teleconnection Pattern Related with the Development of the Okhotsk High and the Northward Progress of the Subtropical High in East Asian Summer, ADVANCES IN ATMOSPHERIC SCIENCES, 20, 237-244.  doi: 10.1007/s00376-003-0009-4
    [12] N. R. Parija, S. K. Dash, 1995: Some Aspects of the Characteristics of Monsoon Disturbances Using a Combined Barotropic-Baroclinic Model, ADVANCES IN ATMOSPHERIC SCIENCES, 12, 487-506.  doi: 10.1007/BF02657007
    [13] Xia LIU, Qiang WANG, Mu MU, 2018: Optimal Initial Error Growth in the Prediction of the Kuroshio Large Meander Based on a High-resolution Regional Ocean Model, ADVANCES IN ATMOSPHERIC SCIENCES, 35, 1362-1371.  doi: 10.1007/s00376-018-8003-z
    [14] WANG Qiang, MU Mu, Henk A. DIJKSTRA, 2012: Application of the Conditional Nonlinear Optimal Perturbation Method to the Predictability Study of the Kuroshio Large Meander, ADVANCES IN ATMOSPHERIC SCIENCES, 29, 118-134.  doi: 10.1007/s00376-011-0199-0
    [15] Gao Shouting, 1988: NONLINEAR ROSSBY WAVE INDUCED BY LARGE-SCALE TOPOGRAPHY, ADVANCES IN ATMOSPHERIC SCIENCES, 5, 301-310.  doi: 10.1007/BF02656754
    [16] Luo Dehai, 1999: Nonlinear Three-Wave Interaction among Barotropic Rossby Waves in a Large-scale Forced Barotropic Flow, ADVANCES IN ATMOSPHERIC SCIENCES, 16, 451-466.  doi: 10.1007/s00376-999-0023-2
    [17] Yaokun LI, Jiping CHAO, Yanyan KANG, 2021: Variations in Wave Energy and Amplitudes along the Energy Dispersion Paths of Nonstationary Barotropic Rossby Waves, ADVANCES IN ATMOSPHERIC SCIENCES, 38, 49-64.  doi: 10.1007/s00376-020-0084-9
    [18] Yaokun LI, Jiping CHAO, Yanyan KANG, 2022: Variations in Amplitudes and Wave Energy along the Energy Dispersion Paths for Rossby Waves in the Quasigeostrophic Barotropic Model, ADVANCES IN ATMOSPHERIC SCIENCES, 39, 876-888.  doi: 10.1007/s00376-021-1244-2
    [19] Ren Shuzhan, 1991: New Approach to Study the Evolution of Rossby Wave Packet, ADVANCES IN ATMOSPHERIC SCIENCES, 8, 79-86.  doi: 10.1007/BF02657366
    [20] Luo Dehai, Ji Liren, 1988: ALGEBRAIC ROSSBY SOLITARY WAVE AND BLOCKING IN THE ATMOSPHERE, ADVANCES IN ATMOSPHERIC SCIENCES, 5, 445-454.  doi: 10.1007/BF02656790

Get Citation+

Export:  

Share Article

Manuscript History

Manuscript received: 16 November 2013
Manuscript revised: 14 January 2014
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

High-Resolution Modeling Study of the Kuroshio Path Variations South of Japan

    Corresponding author: LI Rui; 
  • 1. Physical Oceanography Laboratory/Qingdao Collaborative Innovation Center of Marine Science and Technology, Ocean University of China, Qingdao 266100
Fund Project:  This work was supported by the National Major Research Plan of Global Change (Grant No. 2013CB956201) and the National Natural Science Foundation of China Key Research Project (Grant No. 41130859). Discussions with SUN Shantong and CHEN Zhaohui are greatly appreciated. We are also thankful to the two anonymous reviewers for their constructive comments on the earlier version of the manuscript.

Abstract: A high-resolution ocean general circulation model (OGCM) is used to investigate the Kuroshio path variations south of Japan. The model reproduces many important features of the Kuroshio system including its interannual bimodal variability south of Japan. A decreasing trend of the spatial averaged relative vorticity is detected when the Kuroshio takes the non-large meander (NLM) path, and during the transition period from the NLM to the large meander (LM), a sudden release of velocity shear corresponds well to the weakening of the Shikoku recirculation gyre (SRG), which plays a key role in modulating the Kuroshio path variations. Analysis of eddy energetics indicates that baroclinic instability is mainly responsible for the formation of the LM. In addition, further analysis shows that the strength of the SRG could be largely influenced by the baroclinic Rossby wave adjustment process, forced by the wind stress curl anomalies in the North Pacific basin, based on the model investigation. It is suggested that the cyclonic disturbances might account for the weakening of the SRG, and act as a remote trigger for the baroclinic instability of the Kuroshio south of Japan.

摘要: A high-resolution ocean general circulation model (OGCM) is used to investigate the Kuroshio path variations south of Japan. The model reproduces many important features of the Kuroshio system including its interannual bimodal variability south of Japan. A decreasing trend of the spatial averaged relative vorticity is detected when the Kuroshio takes the non-large meander (NLM) path, and during the transition period from the NLM to the large meander (LM), a sudden release of velocity shear corresponds well to the weakening of the Shikoku recirculation gyre (SRG), which plays a key role in modulating the Kuroshio path variations. Analysis of eddy energetics indicates that baroclinic instability is mainly responsible for the formation of the LM. In addition, further analysis shows that the strength of the SRG could be largely influenced by the baroclinic Rossby wave adjustment process, forced by the wind stress curl anomalies in the North Pacific basin, based on the model investigation. It is suggested that the cyclonic disturbances might account for the weakening of the SRG, and act as a remote trigger for the baroclinic instability of the Kuroshio south of Japan.

1. Introduction
  • The Kuroshio Current, characterized by high eddy activity, is the western boundary current of the North Pacific subtropical gyre. After coming through Tokara Strait (see Fig. 1) from the East China Sea (ECS), the Kuroshio joins the Ryukyu Current, which flows northward along the east side of the Ryukyu Islands, and heads eastward along the southern coast of Japan.

    Figure 1.  Typical paths of the Kuroshio south of Japan (LM, large meander; nNLM, nearshore nonlarge meander; oNLM, offshore nonlarge meander). Solid and dashed lines denote the axes of Kuroshio pathways in AVISO altimetry data (with LM in Jan 2005, nNLM in Nov 2005, and oNLM in Sep 2007) and the model output (with LM in Sep of the model year 28, nNLM in Nov of the model year 34, and oNLM in Apr of the model year 34), respectively, which are both defined as the 110-cm sea SSH isoline. Line PN, OK and ASUKA denote the repeat hydrographic sections.

    A unique feature of the Kuroshio south of Japan is its well-known bimodal path fluctuations, switching between a meandering and a relatively straight one (Fig. 1); namely, the large meander (LM) and non-large meander (NLM) paths (e.g., Taft, 1972). (Kawabe, 1986) further divided the NLM path into two categories——the nearshore non-large meander (nNLM) path and the offshore non-large meander (oNLM) path, with the latter looping southward over the Izu Ridge (near 140°E——see Fig. 1). Both the LM and the NLM states can persist long-range from a few years to a decade, while the transition process between the two paths takes place in only a few months (Kawabe, 1986). Owing to its peculiar oceanographic features and interannual variability, the Kuroshio path variation south of Japan has been the subject of extensive interest for the past several decades.

    Previous studies have suggested the bimodal feature of the Kuroshio south of Japan is a forced phenomenon that depends sensitively on the upstream transport of Kuroshio (e.g., Kawabe, 1980; Chao and McCreary, 1982; Yoon and Yasuda, 1987; Kawabe, 1995; Akitomo et al., 1996). According to a theoretical study (White and McCreary, 1976), the LM can be understood as a stationary Rossby lee wave, and the wavelength is scaled by , where U is the velocity of the Kuroshio and β is the meridional gradient of the Coriolis parameter. This result shows that the zonal scale of the LM could be related to the velocity (or transport) of the upstream Kuroshio. Using sea level data from tide gauges, (Kawabe, 1995) suggested that the LM would occur only when the upstream Kuroshio has large or medium transport, and the NLM path will dominate when the transport is less than 23.5 Sv (106 m3 s-1). However, in contrast, (Qiu and Miao, 2000) suggested that the path oscillation since 1975 could be explained by a self-sustained intrinsic oscillation system involving the Kuroshio and its southern recirculation, based on a two-layer primitive-equation model.

    Recently, high-resolution ocean general circulation model (OGCM) simulations have enabled us to investigate the variability of the Kuroshio path in detail either through case studies (e.g., Endoh and Hibiya, 2001; Miyazawa et al., 2004; Tsujino et al., 2006; Miyazawa et al., 2008; Usui et al., 2008; Endoh and Hibiya, 2009; Usui et al., 2011) or long-term analysis (e.g., Douglass et al., 2012; (Tsujino et al., 2013)). The local dynamics of formation of the LM are discussed extensively in these studies, including the formation and propagation of a trigger meander (e.g., Endoh and Hibiya, 2001; Usui et al., 2008), the effect of bottom topography and deep anticyclone eddies (e.g., Tsujino et al., 2006; Douglass et al., 2012), the role of eddy-Kuroshio interaction (e.g., Miyazawa et al., 2004), or combinations of these factors. In addition to the local dynamics, (Tsujino et al., 2013) revealed the effects of large-scale wind forcing on the bimodality of the Kuroshio path by changing the Kuroshio Current transport, and highlighted the influences of eddies from the recirculation gyre.

    Although both local and remote dynamics have been investigated in previous studies, the mechanisms governing the bimodality of the Kuroshio path still remain elusive due to limited observations and a lack of comparative high-resolution OGCM studies. The purpose of this study is to re-examine the possible mechanisms involved in the formation of the LM, using a (1/10)° OGCM. The model reasonably reproduces the typical paths of the Kuroshio south of Japan, as observed in the altimeter, as well as the interannual variability from the NLM to the LM. Both the local dynamics and remote influence are examined.

    The remainder of the paper is organized as follows. The model used in the study is described in section 2. Section 3 presents the results of the OGCM, and both the mean state and the bimodal variability of the Kuroshio system south of Japan are discussed. Possible mechanisms involved in the formation of the LM are examined in section 4. Finally, section 5 is devoted to a summary and conclusions.

2. Model description
  • The eddy-resolving model used for this study is a version of the Bryan-Cox-Semtner model (Bryan, 1969; Semtner, 1974; Cox, 1984) developed at the University of Wisconsin-Madison (Jacob, 1997; Tobis et al., 1997). The dynamics of this model are based on the Geophysical Fluid Dynamics Laboratory (GFDL) Modular Ocean Model Version 2 (MOM2) (Pacanowski, 1995), with the split-explicit free surface formulation (Killworth et al., 1991). A model performance speed-up is achieved by implementing the gravity wave retardation method with little impact on the internal dynamics (Jensen, 1996; Tobis, 1996). The bathymetry is generated by interpolating the 2-Minute Gridded Global Relief Data (ETOPO2) database. We follow the protocol used to force ocean climate models proposed by (Griffies et al., 2009), and the model is initialized from Polar Science Center Hydrographic Climatology (PHC) with zero velocity. Within three degrees on the northern and southern boundaries of the model domain, the temperature and salinity are restored to the seasonal climatology of PHC 3.0, with the restoring timescale increasing inward linearly from 1 day to 720 days (Marchesiello et al., 2001).

    The quasi-global configuration of the model covers a region from 75°S to 75°N, with a 0.1° grid spacing able to reproduce oceanic mesoscale features such as eddies. The model has 50 vertical levels, and a water depth from 5 m to 5500 m.

    The surface fluxes of wind stress, heat and freshwater used to drive the model are calculated from the Coordinated Ocean-ice Reference Experiment forcing version 2 (CORE.v2; Large and Yeager, 2009) according to the bulk formulae presented by (Large and Yeager, 2004). The model is run globally for 39 years due to a limitation of available computer time. Year 1-25 is considered as the spinup period of the model, and years 26-39 are forced by monthly historical forcing from 1950 to 1963. It should be noted that we focus on the physical mechanisms of the LM in the model rather than a thorough comparison with observations.

3. Model results
  • Figure 2a shows the mean sea surface height (SSH) field of the model during the simulation period. The model succeeds in reproducing the spatial pattern of the Kuroshio and Kuroshio Extension (KE) system, with a reasonable recirculation gyre in the Shikoku Basin and two quasi-stationary meanders [with ridges located at 144° and 150°E, as indicated by (Qiu and Chen, 2005)] after separating from the Japan coast at around 36°N. The annual mean transport of the modeled Kuroshio (above 700 m) across the PN line (Fig. 1) and that of the Ryukyu Current (above 1000 m) across the OK line are about 23.5 and 16.0 Sv, respectively, which are comparable to previous observations (Kawabe, 1995; Andres et al., 2008). Across the ASUKA (Affiliated Surveys of the Kuroshio off Cape Ashizuri) line south of Japan (Fig. 1), the volume transport of the Kuroshio eastward current has been estimated as ranging from 29.0 to 40.2 Sv among three synoptic hydrographic surveys (Nagano et al., 2010). In the present high-resolution model, the annual mean transport across the ASUKA line is 38.5 Sv, which is also consistent with the observed value. In this sense, the present model can resolve the Kuroshio system by and large.

    Figure 2.  Mean SSH (units: cm) for (a) the model, and (b) AVISO. Contour interval is 10 cm. Thick black lines denote the 110-cm SSH isoline, which is regarded as the Kuroshio axis. The white box in (a) indicates the averaged region for relative vorticity.

  • The Kuroshio path fluctuates significantly downstream of the Tokara Strait. A quasi-periodic meandering state has been described using sea level derived from tide gauges since the 1960s (Kawabe, 1985). As shown in Fig. 1, three typical paths of the Kuroshio south of Japan in the model (dashed line) are presented in comparison with that observed by altimeter data (solid line). The altimeter products are collected by the Data Unification and Altimeter Combination System (DUACS) and distributed by Archiving Validation and Interpretation of Satellite Data in Oceanography (AVISO)/the Centre National d'Etudes Spatiales of France. For both the altimeter data and the model output, the Kuroshio axis south of Japan is defined as the 110-cm SSH isoline. As indicated by the thick line in Fig. 2a, the 110-cm SSH contour is consistently located at the maxima of the meridional gradient of the SSH, and can be recognized as a good indicator for the Kuroshio axis (Qiu and Chen, 2005).

    It should be noted that for the past two decades during which the altimeter data are available, only one stable LM case lasting for about 1 year has been observed, with its first appearance in July 2004 [for details, readers are referred to (Miyazawa et al., 2008) and Usui et al. (2008, 2011)], which seems to be much shorter and weaker than historical LM cases (e.g., Kawabe, 1995). In the present model, like most reproductions of the LM in other eddy-resolving models (e.g., Maltrud and McClean, 2005; Douglass et al., 2012; Tsujino et al., 2013), the spatial scale of the LM is much larger than the 2004-05 LM case (see Fig. 1).

    Figure 3.  Yearly paths of the Kuroshio defined by the 110-cm contours in the monthly SSH fields during the simulation period.

    It is noted that the shape of the LM evolves progressively over time along the Kuroshio axis. For a better view of the simulated LM, we plot the yearly paths of the Kuroshio south of Japan in Fig. 3. At the stage of formation (year 28), the initial shape of the LM looks like that observed in the altimeter data, which would not last for the whole lifetime of the LM. After that (year 30), the modeled LM separates from the coastline of Japan at the Tokara Strait south of Kyushu. This process has also been reported by many eddy-resolving models (e.g., Maltrud and McClean, 2005; Douglass et al., 2012; Tsujino et al., 2013), and (Tsujino et al., 2006) stated that this behavior is unrealistic and should be attributed to model bias. In their recent work, (Tsujino et al., 2013) suggested that this can be solved by considering the wind speed as relative to the surface current in evaluating surface stress, which will reduce the model eddy kinetic energy (EKE). Since the observed 2004-05 LM case only lasted for about 1 year, it is still unknown whether this phenomenon would happen in a realistic longer-lived LM.

    Regarding the time scale of the Kuroshio path variations, it is worth mentioning that most previous eddy-resolving models are unable to simulate the Kuroshio path variations with realistic time scale even under realistic historical forcing (e.g., Miyazawa et al., 2004). Using sea level data from tide gauges, (Kawabe, 1995) reported five LM cases with irregular interannual variations during the 1975-91 LM-dominant period. (Qiu and Joyce, 1992) stated that the persistence of the LM varies from 1 year to 5 or 6 years. In subsequent work, (Qiu and Miao, 2000) modeled the interannual variations of the path oscillation with a preferred timescale of 4 years. In the present model, as shown in Fig. 3, during the 14-yr simulated period, the model reproduces the transition from the NLM to the LM three times with a primary LM case developing in year 28 and lasting for about 6 years (LM1). This time scale is comparable to one significant historical LM case which occurred during the period 1975-80 (Kawabe, 1995). This 6-year-long LM ceases at the beginning of year 34 and the Kuroshio stays in a straight state for nearly 3 years before the formation of the second LM. Therefore, the modeled Kuroshio path variability shows evident interannual changes and it would be meaningful to look into the lifecycle of the simulated 6-year-long LM case.

    A general criterion for evaluating the Kuroshio path variations south of Japan is the path length. When the Kuroshio is in its LM state, the path length is much longer than that of the NLM path. Figure 4 shows the path length between 131°E and 141°E during the modeled time, with three LM periods indicated by the dashed black lines. We can see that the simulated 14 years are nearly dominated by the LM state, which explains why the mean Kuroshio shown in Fig. 2a takes the meandering path. To show how the modeled LM evolves, the lifecycle of the meander event beginning in year 27 is shown by a sequence of SSH fields (Fig. 5). Figure 5a shows the Kuroshio in its oNLM path state, with the main stream looping southward over the Izu Ridge. Figure 5b shows the Kuroshio becomes meridionally elongated prior to the formation of the LM. Then, the northward stream steadily switches to the western flank of the Izu Ridge and the LM emerges (Figs. 5c-e). Unlike the observed 2004-05 LM, which developed from a small meander south of Kyushu (e.g., Miyazawa et al., 2008; Usui et al., 2008), the present model presents a similar formation process to that simulated by (Qiu and Miao, 2000). The 6-year-long LM decays together with the southwestward movement of the southern anticyclonic eddy (Figs. 5f and g) and then the Kuroshio path returns to its typical NLM state (Fig. 5h).

    Figure 4.  Time series of the simulated Kuroshio path length south of Japan. The path length is defined as the distance between 131°E and 141°E along the Kuroshio axis. The thick line is smoothed over 1 year (12-month running mean) and the dashed lines below denote the LM periods.

    Figure 5.  Lifecycle of the meander event that developed in year 27 indicated by the sequence of the modeled SSH field. The contour interval is 10 cm and regions with SSH <110 cm are shaded.

    Although the overall performance of the simulation seems to be good, especially in the region south of Japan, discrepancies between the model and the observations nevertheless exist. As shown in Fig. 2a, after separating from the Japan coast at around 36°N, the mean upstream KE jet develops a larger-amplitude meander than that observed in the altimeter data (Fig. 2b). With a closer look at the yearly path of the Kuroshio shown in Fig. 3, we find that when the Kuroshio is in its LM state, the upstream KE jet tends to shift northward more often than that of NLM state. Using SSH data from multiple satellite altimeters, (Qiu and Chen, 2005) pointed out that the meridional path change of the KE jet is independent of the bimodal Kuroshio path variability south of Japan. Therefore, it seems that the present model may not show good skill in simulating the amplitude of the upstream KE jet migration, but this aspect is beyond the scope of the following analysis. Another shortcoming is that the decaying process of the 6-year-long LM is not an observed feature of the 2004-05 LM case (e.g., Usui et al., 2011). Although a significant impact of anticyclonic eddies on the formation of the LM has been clarified by previous studies (e.g., Mitsudera et al., 2001; Waseda et al., 2003), the present model indicates that movement of anticyclonic eddies might also play a role in weakening the meander. In this case, the focus of our study is on identifying possible conditions for the development of the LM, more than on its recession. Despite the problems with our model results, we believe that the present model captures the essential characteristics of the Kuroshio system and that the Kuroshio path variability is adequately simulated in the region south of Japan.

4. Dynamics for the formation of the LM
  • In the self-sustained oscillation system proposed by (Qiu and Miao, 2000), they concluded that the accumulation of low-PV (potential vorticity) anomalies intensifies the offshore recirculation gyre south of Japan and forces the Kuroshio to flow along the Japan coast when the Kuroshio is in its straight path state. And this process could increase the vertical velocity shear and further lead to baroclinic instability of the straight-path Kuroshio, which causes the development of the LM.

    In the present study, following (Qiu and Miao, 2000), the relative vorticity, ζ=(? v/? x-? u/? y), is averaged in the Kuroshio region of (25°-35°N, 132°-140°E) (boxed region in Fig. 2a), where the horizontal and meridional velocity for the calculation is averaged vertically over the upper 500 m. As shown in Fig. 6a, the relative vorticity is calculated over the above area and a decreasing trend is found when the Kuroshio is in its straight path state, indicating the accumulation of low-PV anomalies south of Japan. While in the LM state, the area-averaged ζ is relatively stable due to the presence of a cyclonic eddy north of the Kuroshio axis.

    Following the idea of the self-sustained oscillation mechanism, we then calculate the strength of the Shikoku recirculation gyre (SRG) and the vertical shear of the alongshore Kuroshio, the results of which are shown in Figs. 6b and c, respectively. Here, we should distinguish the SRG from the anticyclonic eddy south of the Kuroshio. The SRG is associated with the basin-scale circulation and is much larger than the anticyclonic eddy (≈200 km in diameter for the velocity core) (Mitsudera et al., 2001). To evaluate the strength of the SRG, we follow (Qiu and Chen, 2005) to define:

    where A denotes the area of SSH >150 cm in the boxed region, where the relative vorticity is calculated (Fig. 2a). As shown in Fig. 2a, the 150-cm isoline matches the outer boundaries of the SRG well. Besides, the vertical shear of the alongshore Kuroshio is defined as U0-200-U200-600, where U0-200 and U200-600 denote the vertically-averaged alongshore Kuroshio velocity over the upper 200 m and 200-600 m, respectively. The SRG strength increases continuously when the Kuroshio is in its straight path state, corresponding to the decreasing trend of the averaged relative vorticity. While the increasing trend of the vertical shear of mean velocity is not obtained as in (Qiu and Miao, 2000) when the Kuroshio is in NLM state, the occurrences of the LMs do accompany a sharp decrease in the velocity shear, which agrees with their results.

    By comparing the SRG strength with velocity shear (Fig. 6), at the formation of the LMs, it is readily concluded that the sudden release of velocity shear corresponds well with the weakening of the SRG. It seems that the weakening of the recirculation gyre could help the fully-charged velocity shear to break down; that is, through baroclinic instability.

    Figure 6.  (a) Time series of relative vorticity, ζ, averaged in the Kuroshio region of (25°-35°N, 132°-140°E), where horizontal and meridional velocity for the calculation is vertically averaged in the upper 500 m. (b) Time series of the southern recirculation strength in the region same as ζ. (c) Time series of the vertical shear of the alongshore Kuroshio, U0-200-U200-600, where U0-200 and U200-600 denote the vertically averaged alongshore Kuroshio velocity of the upper 200 m and 200-600 m, respectively; denotes the spatial average in the region (30°-35°N, 132°-140°E). For all three time series, the thick black lines are smoothed over 12 months. LM periods are shaded in the three panels.

  • Previous studies have emphasized the importance of baroclinic instability on the formation of the LM (e.g., Yoon and Yasuda, 1987; Hurlburt et al., 1996; Endoh and Hibiya, 2001; Tsujino et al., 2006). A useful tool to evaluate the flow instabilities is eddy energetics analysis. However, due to the great fluctuations of the Kuroshio path south of Japan, it is not easy to define the background flow field in a satisfactory way. (Tsujino et al., 2006) proposed decomposing the instantaneous field into the field at the previous time interval (mean field) and the deviation from it (eddy field). Here, we adopt the same method in the monthly model outputs and analyze the local energetics.

    Following (Tsujino et al., 2006), the barotropic (BT) conversion rate from the mean kinetic energy (MKE) to eddy kinetic energy (EKE), indicating the barotropic instability, is defined as

    where ρ0 is the typical value of the seawater density under the Boussinesq approximation, is the eastward (northward) velocity in the mean field, and u'(v') is the eastward (northward) perturbed velocity in the eddy field.

    The baroclinic (BC) conversion rate from mean potential energy (MPE) to eddy potential energy (EPE), indicating the baroclinic instability, is defined as:

    where g is the gravity acceleration, u'=(u',v',0) is the perturbed field of a geostrophic flow, is a horizontal gradient operator, is the depth-dependent background density field, and δρ(x,y,z,t) is the weakly varying density field, with the overbar and prime denoting the mean and eddy component, respectively.

    Besides, considering that the rotational component of the eddy buoyancy flux () plays no role in converting energy between mean and eddy fields (Tsujino et al., 2006), the dynamically baroclinic (dynBC) conversion rate, which only takes the divergent component of the eddy buoyancy flux in the BC conversion rate, is also calculated due to its dynamical importance:

    For a detailed discussion on the calculation of energy conversion rates, readers are referred to Tsujino et al. (2006, Appendix B).

    The three conversion rates (200 m) at the formation of the modeled 6-year-long LM are compared in Fig. 7. It is seen that both the BT and the BC conversion rates show high positive values when the northward Kuroshio stream switches to the western flank of Izu Ridge and the LM emerges, indicating the important impact of flow instabilities on the formation process. As expected, both the BC and the dynBC show a much more significant signal than the BT, suggesting that the baroclinic instability bears greater responsibility for the growth of the LM than the barotropic instability. Another feature is that the large positive values of the BC and the dynBC are located between the Japan coast and the Kuroshio axis, where a cyclonic eddy accompanying the LM forms.

    Figure 7.  Energy conversion rates (shading; units: kg m-1 s-3) at 200 m on the formation of the 6-year-long LM for (a) BT, (b) BC, and (c) dynBC. Contours denote the potential density at 200 m (contour interval = 0.15 kg m-3).

    In this case, we calculate the volume-integrated energy conversion rates over the area of SSH <40 cm in the boxed Kuroshio region of (25°-35°N, 132°-140°E) (Fig. 8). A sharp increase in all these three conversion rates can be seen before the onset of the LM, indicating energy transfer from the mean to the eddy field. Comparing with the vertical shear of the alongshore Kuroshio shown in Fig. 6c, we find that such an increase in conversion rates corresponds well to the sudden breakdown of the velocity shear. By averaging the volume-integrated conversion rates over five months before the onset of each LM event (two months for LM3), it is shown that both the BC and the dynBC are at least twice as much as the BT (Table 1). Thus, the above analyses confirm that baroclinic instability is mainly responsible for the formation of the LM, which agrees with previous studies (e.g., Endoh and Hibiya, 2001; Miyazawa et al., 2004).

    Figure 8.  Time series of three conversion rates volume-integrated (upper 1000 m) over the area where SSH <40 cm in the boxed Kuroshio region of (25°-35°N, 132°-140°E). LM periods are shaded.

  • Relations between the upstream transport and the bimodality south of Japan have been discussed extensively, based on both observational data (e.g., Kawabe, 1980, 1995; Qiu and Miao, 2000) and modeling studies (e.g., Chao and McCreary, 1982; Yoon and Yasuda, 1987; Akitomo et al., 1996; Douglass et al., 2012). Here, we also re-examine the possible interaction between the upstream transport and the path variability.

    Figure 9 shows the volume transport across the Tokara Strait during the 14-yr period. Although each of the three LM cases accompanies a transport higher than 23.5 Sv (Kawabe, 1995), the transport decreases several months before the LM forms when the meander gradually develops, indicating that the generation of the LM may not be explained by the increase in upstream transport. However, by comparing the SRG strength to the transport through Tokara Strait, we find an interesting correlation between them (1-yr smoothed, lagged correlation = 0.65), with the recirculation strength leading the transport by 7 months. Such correspondence is much more significant when the Kuroshio takes the meandering path and the SRG stays on the southwest of the main body of the LM, suggesting an impact of downstream (south of Japan) path variability on the temporal change of upstream (ECS) volume transport. Recently, using the Parallel Ocean Program (POP), (Douglass et al., 2012) also showed a similar phenomenon and suggested that the increase of transport through the Tokara Strait might be a response to the shift of the LM state. They stated that during the LM, the SRG could force the transport through the Kerama Gap (shown in Fig. 1) and further increase the Kuroshio transport in ECS, owing to its southwestward position compared with that in the NLM state.

    Figure 9.  Time series of the SRG strength and volume transport through the Tokara Strait.Thick gray lines denote the LM periods.

    Considering that the SRG also plays an essential role in modulating the Kuroshio path variability south of Japan, it seems that the present model indicates a regional oscillation system that includes the SRG, volume transport of the Kuroshio in ECS and the Kuroshio path bimodality south of Japan. The potential interactions among the three elements involved in the regional system are worth considering for future studies.

  • The above analysis points to the important role of the SRG strength in modulating the regional transport and path variations of the Kuroshio. But what causes the temporal change of the SRG strength? As an inseparable part of the Kuroshio system, the SRG arises from the accumulation of low-PV anomalies carried northward by the western boundary current, indicated by both theoretical and modeling studies (e.g., Cessi et al., 1987; Ierley and Young, 1988; Liu, 1997).

    To determine what controls the temporal variation of SRG, we show the sea surface height anomaly (SSHA) along 29°N in Fig. 10b, with the dashed black line denoting the westward propagating speed of 0.048 m s-1, which is comparable to the phase speed of first-mode baroclinic Rossby waves estimated by the T/P SSH data (Qiu, 2003). Large anticyclonic disturbances with positive SSHA can be found before the formation of the LMs (green lines in Fig. 10b). These positive SSHAs can be traced back to the eastern North Pacific basin (160°-140°W), and were presumably induced by negative wind stress curl anomalies in the central to eastern North Pacific basin, propagating westward as baroclinic Rossby waves. By comparing with the SRG strength shown in Fig. 10a, we find that these westward-propagating disturbances correlate well with the variability of the SRG strength. During the NLM state, the anticyclonic disturbances with low-PV water accumulate west of the Izu Ridge (near 140°E, see Fig. 1), strengthening the offshore SRG. Onsets of the LMs are accompanied by a shift from anticyclonic disturbances to cyclonic ones. As discussed in section 4.1, weakening of the recirculation gyre would help the velocity shear to break down and cause the LM. In this case, the westward-propagating negative SSHAs might account for the weakening of the SRG before the formation of the LMs, and act as a remote trigger for the baroclinic instability of the Kuroshio south of Japan. It is worth mentioning that for the observed 2004-05 LM event, (Usui et al., 2008) also reported remarkable mid-latitude westward-propagating negative SSHAs that contribute substantially to the formation of the LM.

    Figure 10.  (a) Time series of the 1-yr smoothed SRG strength (RS); for comparison with (b), it is plotted as the deviation from the mean RS value. (b) Hovmöller diagram of SSH anomaly along 29°N. The dashed black line denotes the westward-propagating speed of 0.048 m s-1, and green lines denote the LM periods.

    (Tsujino et al., 2013) also mentioned the remote forcing conditions as potential factors that can affect the formation of the LM. They suggested that the wind-induced low-latitude positive SSHA could be advected by the Kuroshio in the ECS, causing an increase of the Kuroshio transport, which promotes the formation of the LM. In comparison, our analysis shows that, being independent of the upstream Kuroshio transport, the remote disturbances caused by large-scale wind in the mid-latitudes may directly influence the SRG strength, stimulating the variability of the Kuroshio path south of Japan.

5. Conclusion
  • In this study, a high-resolution OGCM with realistic bathymetry was used to model the Kuroshio path variations south of Japan. It succeeds in reproducing many important aspects of the Kuroshio system with its interannual bimodal variability south of Japan. During the 14-yr simulation, the model reproduces the transition from the NLM to LM three times with a primary LM that lasts for about 6 years, which is comparable to an historical LM case (1975-80). Possible mechanisms for the development of the LM have been examined in this paper.

    Following the idea of the self-sustained oscillation mechanism, we first examined the Kuroshio path variability associated with the local relative vorticity, the SRG strength, and the vertical velocity shear of the alongshore Kuroshio. The spatial averaged relative vorticity shows a decreasing trend when the Kuroshio takes the NLM path, indicating the accumulation of low-PV anomalies south of Japan. During the transition period from the NLM to LM, a sudden release of velocity shear corresponds well to the weakening of the SRG, which plays a key role in modulating the Kuroshio path variations. Analysis of eddy energetics indicated that both the BC and the dynBC show a much more significant signal than the BT, suggesting that baroclinic instability bears greater responsibility for the growth of the LM than the barotropic instability. A sharp increase in the volume-integrated conversion rates can be seen before the onset of the LM, which also corresponds well to the sudden breakdown of the alongshore velocity shear. This study confirms the previous conclusion that baroclinic instability is mainly responsible for the formation of the LM.

    Relations between the upstream transport and the Kuroshio path variability south of Japan have also been examined. Despite the temporal limit of the present study, an interesting connection was found between the SRG strength and the transport through the Tokara Strait, with the recirculation strength leading the transport by 7 months.

    The simulation also indicated a possible mid-latitude remote forcing that could directly influence the strength of the SRG. The large anticyclonic (cyclonic) disturbances with positive (negative) SSHAs, propagating westward as baroclinic Rossby waves adjustment to the wind stress curl anomalies in the North Pacific basin, correlate well with the variability of the SRG strength. In addition, onsets of the LMs accompany the switch from anticyclonic disturbances to cyclonic ones. Thus, we propose that the negative SSHAs might account for the weakening of the SRG before the formation of the LMs, and act as a remote trigger for the baroclinic instability of the Kuroshio.

    Overall, the present model under realistic historical wind forcing reproduces many important features of the Kuroshio system. In the region south of Japan, our analysis emphasizes both local and remote roles in modulating the flow instability and regional oceanic variability. Although disparities exist between the modeled LM and the 2004-05 LM case, the present study might provide some insight into the intrinsic oceanographic mechanisms and the effect of large-scale wind forcing to better understand the Kuroshio variability south of Japan.

Reference

Catalog

    /

    DownLoad:  Full-Size Img  PowerPoint