Skip to main content
Log in

Effects of large-scale wind on the Kuroshio path south of Japan in a 60-year historical OGCM simulation

  • Published:
Climate Dynamics Aims and scope Submit manuscript

A Correction to this article was published on 11 September 2018

Abstract

The effects of large-scale wind forcing on the bimodality of the Kuroshio path south of Japan, the large meander (LM) and non-large meander (NLM), were studied by using a historical simulation (1948–2007) with a high-resolution Ocean general circulation models (OGCM). The Kuroshio in this simulation spent much time in the NLM state, and reproduced several aspects of its long-term path variability for the first time in historical OGCM simulation, presumably because the eddy kinetic energy was kept at a moderate level. By using the simulated fields, the relationships between wind forcing (or Kuroshio transport) and path variation proposed by past studies were tested, and specific roles of eddies in those variations were investigated. The long-term variation of the simulated net Kuroshio transport south of Japan was largely explained by the linear baroclinic Rossby wave adjustment to wind forcing. In the simulated LM events, a triggering meander originated from the interaction of a wind-induced positive sea surface height (SSH) anomaly with the upstream Kuroshio and was enlarged by cyclonic eddies from the recirculation gyre. The cyclonic eddy of the trigger meander was followed by a sizable anticyclonic eddy on the upstream side. Subsequently, a weak (strong) Kuroshio favored the LM (NLM). The LM tended to be maintained when the Kuroshio transport off southern Japan was small, and increasing Kuroshio transport promoted decay of an existing LM. The supply of disturbances from upstream, which is related to the wind-induced SSH variability at low latitudes, contributed to the maintenance of an existing LM.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11
Fig. 12
Fig. 13
Fig. 14
Fig. 15
Fig. 16
Fig. 17
Fig. 18

Similar content being viewed by others

References

  • Akitomo K (2008) Effects of stratification and mesoscale eddies on Kuroshio path variation south of Japan. Deep Sea Res I 55:997–1008. doi:10.1016/j.dsr.2008.03.013

    Article  Google Scholar 

  • Akitomo K, Ooi M, Awaji T, Kutsuwada K (1996) Interannual variability of the Kuroshio transport in response to wind stress field of the North Pacific: its relation to the path variation south of Japan. J Geophys Res 101:14057–14071

    Article  Google Scholar 

  • Bartlett MS (1935) Some aspects of the time-correlation problem in regard to tests of significance. J R Stat Soc 98:536–543

    Article  Google Scholar 

  • Bretherton CS, Widmann M, Dymnikov VP, Wallace JM, Blade I (1999) The effective number of spatial degree of freedom of a time-varying field. J Clim 12:1990–2009

    Article  Google Scholar 

  • Collecte, Localisation, Satellites (CLS) (2010) SSALTO/DUACS user handbook: (M)SLA and (M)ADT near-real time and delayed time products. Rep CLS-DOS-NT-06.304, Centre Nat d’Etud Spatiales, Toulouse, France, p 44

  • Cushman-Roisin B (1993) Trajectories in Gulf stream meanders. J Geophys Res 98:2543–2554

    Article  Google Scholar 

  • Endoh T, Hibiya T (2001) Numerical study of the generation and propagation of trigger meander of the Kuroshio south of Japan. J Geophys Res 106:26833–26850

    Article  Google Scholar 

  • Fujii Y, Tsujino H, Usui N, Nakano H, Kamachi M (2008) Application of singular vector analysis to the Kuroshio large meander. J Geophys Res 113:C07026. doi:10.1029/2007JC004476

    Article  Google Scholar 

  • Gent PR, McWilliams JC (1990) Isopycnal mixing in ocean circulation models. J Phys Oceanogr 20:150–155

    Article  Google Scholar 

  • Griffies SM, Hallberg RW (2000) Biharmonic friction with a Smagorinsky-like viscosity for use in large-scale eddy-permitting ocean models. Mon Wea Rev 128:2935–2946

    Article  Google Scholar 

  • Griffies SM, Biastoch A, Bö ning C, Bryan F, Danabasoglu G, Chassignet EP, England MH, Gerdes R, Haak H, Hallberg RW, Hazeleger W, Jungclaus J, Large WG, Madec G, Pirani A, Samuels BL, Scheinert M, Sen Gupta A, Severijns CA, Simmons HL, Treguier AM, Winton M, Yeager S, Yin J (2009) Coordinated ocean-ice reference experiments (COREs). Ocean Modell 26:1–46. doi:10.1016/j.ocemod.2008.08.007

    Article  Google Scholar 

  • Hasumi H (2006) CCSR ocean component model (COCO) version 4.0. CCSR report, No.25, Univ Tokyo, Tokyo, Japan, p 103

  • Hurlburt HE, Wallcraft AJ, Schmitz WJ Jr, Hogan PJ, Metzger EJ (1996) Dynamics of the Kuroshio/Oyashio current system using eddy-resolving models of the North Pacific Ocean. J Geophys Res 101:941–976

    Article  Google Scholar 

  • Imawaki S, Uchida H, Ichikawa H, Fukawasa M, Umatani S, The ASUKA Group (2001) Satellite altimeter monitoring the Kuroshio transport south of Japan. Geophys Res Lett 28:17–20

    Article  Google Scholar 

  • Isern-Fontanet J, Garcia-Landona E, Font J (2003) Identification of marine eddies from altimetry. J Atmos Oceanic Technol 20:772–778

    Article  Google Scholar 

  • Ishizaki H, Motoi T (1999) Reevaluation of the Takano-Onishi scheme for momentum advection on bottom relief in ocean models. J Atmos Oceanic Technol 16:1994–2010

    Article  Google Scholar 

  • Kalnay E et al (1996) The NCEP/NCAR 40-year reanalysis project. Bull Am Meteor Soc 77:437–471

    Article  Google Scholar 

  • Kawabe M (1986) Transition processes between the three typical paths of the Kuroshio. J Oceanogr Soc Japan 42:174–191

    Article  Google Scholar 

  • 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

    Article  Google Scholar 

  • Killworth PD, Webb DJ, Stainforth D, Paterson SM (1991) The development of a free-surface Bryan-Cox-Semtner ocean model. J Phys Oceanogr 21:1333–1348

    Article  Google Scholar 

  • Killworth PD, Blundell JR (2003) Long extratropical planetary wave propagation in the presence of slowly varying mean flow and bottom topography. Part I: the local problem. J Phys Oceanogr 33:784–801

    Article  Google Scholar 

  • Kurogi M, Akitomo K (2006) Effects of stratification on the stable paths of the Kuroshio and on their variation. Deep Sea Res I 53:1564–1577. doi:10.1016/j.dsr.2006.07.003

    Article  Google Scholar 

  • Large WG, Yeager SG (2004) Diurnal to decadal global forcing for ocean and sea-ice models: the data sets and flux climatologies. Technical Report TN-460+STR, NCAR, p 105

  • Large WG, Yeager SG (2009) The global climatology of an interannually varying air sea flux data set. Clim Dyn 33:341–364. doi:10.1007/s00382-008-0441-3

    Article  Google Scholar 

  • Leonard BP (1979) A stable and accurate convective modeling procedure based upon quadratic upstream interpolation. J Comput Methods Appl Mech Eng 19:59–98

    Article  Google Scholar 

  • Leonard BP, MacVean MK, Lock AP (1993) Positivity-preserving numerical schemes for multidimensional advection. Tech Memo, 106055, ICOMP-93-05, NASA, Washington DC, p 62

  • Mantua NJ, Hare SR, Zhang Y, Wallace JM, Francis RC (1997) A Pacific interdecadal climate oscillation with impacts on salmon production. Bull Am Meteorol Soc 78:1069–1079

    Article  Google Scholar 

  • Masuda A (1982) An interpretation of the bimodal character of the stable Kuroshio path. Deep Sea Res 29:471–484

    Article  Google Scholar 

  • Miyazawa Y, Guo X, Yamagata T (2004) Roles of mesoscale eddies in the Kuroshio paths. J Phys Oceanogr 34:2203–2222

    Article  Google Scholar 

  • Miyazawa Y, Kagimoto T, Guo X, Sakuma H (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

    Article  Google Scholar 

  • Mizuno K, White WB (1983) Annual and interannual variability of the Kuroshio current system. J Phys Oceanogr 13:1847–1867

    Article  Google Scholar 

  • Nakamura H, Nishina A, Minobe S (2012) Response of storm tracks to bimodal Kuroshio path states south of Japan. J Clim 25:7772–7779. doi:10.1175/JCLI-D-12-00326.1

    Article  Google Scholar 

  • Nakano H, Suginohara N (2002) Effects of bottom boundary layer parameterization on reproducing deep and bottom waters in a world ocean model. J Phys Oceanogr 32:1209–1227

    Article  Google Scholar 

  • Nakano H, Tsujino H, Furue R (2008) The Kuroshio current system as a jet and twin “relative” recirculation gyres embedded in the Sverdrup circulation. Dyn Atmos Oceans 45:135–164. doi:10.1016/j.dynatmoce.2007.09.002

    Article  Google Scholar 

  • Nakano H, Ishikawa I (2010) Meridional shift of the Kuroshio Extension induced by response of recirculation gyre to decadal wind variations. Deep Sea Res II 57:1111–1126. doi:10.1016/j.dsr2.2009.12.002

    Article  Google Scholar 

  • Noh Y, Kim H-J (1999) Simulations of temperature and turbulence structure of the oceanic boundary layer with the improved near-surface process. J Geophys Res 104:15621–15634

    Article  Google Scholar 

  • Oka E, Kawabe M (2003) Dynamic structures of the Kuroshio south of Kyushu in relation to the Kuroshio path variations. J Oceanogr 59:595–608

    Article  Google Scholar 

  • Okubo A (1970) Horizontal dispersion of floatable particles in the vicinity of velocity singularities such as convergences. Deep Sea Res 17:445–454

    Google Scholar 

  • Prather MJ (1986) Numerical advection by conservation of second-order moments. J Geophys Res 91:6671–6681

    Article  Google Scholar 

  • Qiu B (2003) Kuroshio extension variability and forcing of the Pacific decadal oscillations: responses and potential feedback. J Phys Oceanogr 33:2465–2482

    Article  Google Scholar 

  • Qiu B, Miao W (2000) Kuroshio path variations south of Japan: bimodality as a self-sustained internal oscillation. J Phys Oceanogr 30:2124–2137

    Article  Google Scholar 

  • Qiu B, Chen S (2005) Variability of the Kuroshio extension jet, recirculation gyre, and mesoscale eddies on decadal timescales. J Phys Oceanogr 35:2090–2103. doi:10.1175/JPO2807.1

    Article  Google Scholar 

  • Qiu B, Chen S (2010) Eddy-mean flow interaction in the decadally modulating Kuroshio extension system. Deep Sea Res II 57:1098–1110. doi:10.1016/j.dsr2.2008.11.036

    Article  Google Scholar 

  • Redi MH (1982) Oceanic isopycnal mixing by coordinate rotation. J Phys Oceanogr 12:1154–1158

    Article  Google Scholar 

  • Sakamoto T, Hasumi H, Ishii M, Emori S, Suzuki T, Nishimura T, Sumi A (2005) Responses of the Kuroshio and the Kuroshio extension to global warming in a high-resolution climate model. Geophys Res Lett 32:L14617. doi:10.1029/2005GL023384.

    Article  Google Scholar 

  • Sasaki H, Xie S-P, Taguchi B, Nonaka M, Masumoto Y (2010) Seasonal variations of the Hawaiian Lee countercurrent induced by the meridional migration of the trade winds. Ocean Dyn 60:705–715. doi:10.1007/s10236-009-0258-6

    Article  Google Scholar 

  • Sato Y, Yukimoto S, Tsujino H, Ishizaki H, Noda A (2006) Response of North Pacific Ocean circulation in a Kuroshio-resolving ocean model to an Arctic oscillation (AO)-like change in northern hemisphere atmospheric circulation due to greenhouse-gas forcing. J Meteorol Soc Japan 84:295–309. doi:10.2151/jmsj.84.295

    Article  Google Scholar 

  • Shoji D (1972) Time variation of the Kuroshio south of Japan. In: Stommel H, Yoshida Y (eds) Kuroshio, physical aspects of the Japan current. University of Tokyo Press, Tokyo, Japan, pp 217–234

    Google Scholar 

  • Smagorinsky J (1963) General circulation experiments with the primitive equations. I. The basic experiment. Mon Wea Rev 91:99–164

    Article  Google Scholar 

  • Smith RD, JC McWilliams (2003) Anisotropic horizontal viscosity for ocean models. Ocean Modell 5:129–156

    Article  Google Scholar 

  • St Laurent LC, Simmons HL, Jayne SR (2002) Estimating tidally driven mixing in the deep ocean. Geophys Res Lett 29:2106–2110. doi:10.1029/2002GL015633

    Article  Google Scholar 

  • Sugimoto S, Hanawa K, Narikiyo K, Fujimori M, Suga T (2010) Temporal variations of the net Kuroshio transport and its relation to atmospheric variations. J Oceanogr 66:611–619

    Article  Google Scholar 

  • Sugimoto S, Hanawa K (2012) Relationship between the path of the Kuroshio in the south of Japan and the path of the Kuroshio extension in the east. J Oceanogr 68:219–225. doi:10.1007/s10872-011-0089-1

    Article  Google Scholar 

  • Taguchi B, Xie S-P, Schneider N, Nonaka M, Sasaki H, Sasai Y (2007) Decadal variability of the Kuroshio extension: observations and an eddy-resolving model hindcast. J Clim 20:2357–2377. doi:10.1175/JCLI4142.1

    Article  Google Scholar 

  • Tsujino H, Usui N, Nakano H (2006) Dynamics of Kuroshio path variations in a high-resolution general circulation model. J Geophys Res 111:C11001. doi:10.1029/2005JC003118

    Article  Google Scholar 

  • Tsujino H, Motoi T, Ishikawa I, Hirabara M, Nakano H, Yamanaka G, Yasuda T, Ishizaki H (2010) Reference manual for the Meteorological Research Institute Community Ocean Model (MRI.COM) Version 3. Technical Reports of the Meteorological Research Institute, No 59, Meteorol Res Inst, Tsukuba, Japan, p 241. (Available at http://www.mri-jma.go.jp/Publish/Technical/DATA/VOL_59/59.html)

  • Tsujino H, Hirabara M, Nakano H, Yasuda T, Motoi T, Yamanaka G (2011) Simulating present climate of the global ocean-ice system using the Meteorological Research Institute Community Ocean Model (MRI. COM): simulation characteristics and variability in the Pacific sector. J Oceanogr 67:449–479. doi:10.1007/s10872-011-0050-3

    Article  Google Scholar 

  • Usui N, Tsujino H, Fujii Y, Kamachi M (2006) Short-range prediction experiments of the Kuroshio path variabilities south of Japan. Ocean Dyn 56:607–623. doi:10.1007/s10236-006-0084-z

    Article  Google Scholar 

  • Usui N, Tsujino H, Fujii Y, Kamachi M (2008) Generation of a trigger meander for the 2004 Kuroshio large meander. J Geophys Res 113:C01012. doi:10.1029/2007JC004266

    Article  Google Scholar 

  • Usui N, Tsujino H, Nakano H, Fujii Y (2008) Formation process of the Kuroshio large meander in 2004. J Geophys Res 113:C08047. doi:10.1029/2007JC004675

    Article  Google Scholar 

  • Usui N, Tsujino H, Nakano H, Fujii Y, Kamachi M (2011) Decay mechanism of the 2004/05 Kuroshio large meander. J Geophys Res 116:C10010. doi:10.1029/2011JC007009

    Article  Google Scholar 

  • Weiss J (1991) The dynamics of enstrophy transfer in two-dimensional hydrodynamics. Physica D 48:273–294

    Article  Google Scholar 

  • White WB, McCreary JP (1976) The Kuroshio meander and its relationship to large-scale circulation. Deep Sea Res 23:33–47

    Google Scholar 

  • Xu H, Tokinaga H, Xie S-P (2010) Atmospheric effects of the Kuroshio large meander during 2004–05. J Clim 23:4704–4715. doi:10.1175/2010JCLI3267.1

    Article  Google Scholar 

  • Yamanaka G, Ishizaki H, Hirabara M, Ishikawa I (2008) Decadal variability of the subtropical front of the western North Pacific in an eddy-resolving ocean general circulation model. J Geophys Res 113:C12027. doi:10.1029/2008JC005002

    Article  Google Scholar 

  • Yasuda T, Sakurai K (2006) Interdecadal variability of the sea surface height around Japan. Geophys Res Lett 33:L01605. doi:10.1029/2005GL024920

    Article  Google Scholar 

  • Yoshimura H, Yukimoto S (2008) Development of a simple coupler (Scup) for earth system modeling. Pap Meteor Geophys 59:19–29

    Article  Google Scholar 

  • Yoshinari H, Ikeda M, Tanaka K, Masumoto Y (2004) Sensitivity of the interannual Kuroshio transport variation south of Japan to wind dataset in OGCM calculation. J Oceanogr 60:341–350

    Article  Google Scholar 

  • Zhai X, Greatbatch RJ (2007) Wind work in a model of the northwest Atlantic Ocean. Geophys Res Lett 34:L04606. doi:10.1029/2006GL028907

    Article  Google Scholar 

Download references

Acknowledgments

We are grateful for constructive comments by Masafumi Kamachi and Yosuke Fujii on the local dynamics of the Kuroshio and by Tamaki Yasuda on the long-term climate variability of the North Pacific Ocean. Comments from anonymous reviewers greatly improved the earlier version of the manuscript. This work was funded by the Meteorological Research Institute and was partly supported by the Japan Society for the Promotion of Science (JSPS) through a Grant-in-Aid for Scientific Research (B) 21340133 and by the Ministry of Education, Culture, Sports, Science and Technology (MEXT) through a Grant-in-Aid for Scientific Research on Innovative Areas 22106006. SN is supported by the research project “KAKUSHIN”, funded by MEXT.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Hiroyuki Tsujino.

Appendices

Appendix 1: Linear baroclinic vorticity model

To understand the long-term variation of the simulated Kuroshio transport, the result is compared with that of the linearized quasi-geostrophic 1.5-layer reduced gravity model that governs the large-scale, baroclinic ocean response to surface wind stress curl forcing (Qiu 2003). The governing equation is

$$ \frac{\partial \Uppsi_{\rm Sv}}{\partial t}-c_R \frac{\partial \Uppsi_{\rm Sv}} {\partial x} = - \frac{g^{\prime}H} {\rho_0 f^2} \hbox{curl} \tau, $$
(2)

where \(\Uppsi_{\rm Sv}\) is the vertically integrated horizontal transport stream function relative to the eastern boundary, c R is the speed of the long baroclinic Rossby waves, \(g^{\prime}\) is the reduced gravity, ρ0 is the reference density, f is the Coriolis parameter, H is the upper layer thickness, and curl τ is the wind stress curl. Integrating (2) from the eastern boundary (x e ) along the baroclinic Rossby wave characteristic, we have

$$ \Uppsi_{\rm Sv} (x,y,t) = \frac{g^{\prime}H} {\rho_0 f^2 c_{R} } \int\limits^{x}_{x_e} \hbox{curl} \tau \left(x^{\prime}, y, t + \frac{x-x^{\prime}} {c_R} \right) dx^{\prime}, $$
(3)

where we take \(\Uppsi_{\rm Sv} (x_e,y,t) = 0\). The layer thickness H is implicitly adjusted so that Eq. (2) becomes the Sverdrup balance in a steady state, that is,

$$ H = \frac{f^2 c_R} {g^{\prime} \beta}, $$
(4)

yielding

$$ \Uppsi_{\rm Sv} (x,y,t) = \frac{1} {\rho_0 \beta} \int\limits^{x} _{x_e} \hbox{curl} \tau \left(x^{\prime}, y, t + \frac{x-x^{\prime}} {c_R} \right) dx^{\prime}, $$
(5)

where β is the meridional gradient of the Coriolis parameter.

To hindcast \(\Uppsi_{\rm Sv}\) using Eq. (5), monthly wind stress data either recalculated during the model run (for REL) or given by the reanalysis (for ABS) is used. The speed of the baroclinic Rossby wave (c R ) is taken from Qiu (2003) for mid-latitudes and from Killworth and Blundel (2003) for low-latitudes.

Appendix 2: Definition of the Kuroshio path

The simulated Kuroshio paths are classified into four states (LM, NNLM, ONLM, and TSLM; Fig. 1) on the basis of the southward extent of water colder than 15°C at 300 m depth from the southern coast of Japan and the longitude of the main body of the meander defined as the location of the minimum barotropic stream function in this cold water region. Note that Mizuno and White (1983) defined the Kuroshio path in the Kuroshio Extension region by the location at 300 m depth of the 12°C isotherm. Our choice of 15°C is made for two reasons: the region of interest is in the warmer upstream region and the model has a slightly warm bias. We used the following procedures:

  • If the maximum southward extent of the cold-water region between 135.5°E and 140.5°E is south (north) of 32.5°N, the Kuroshio is determined to be in the LM (NNLM) state, with the following exceptions:

  • The LM is determined to be the TSLM if its southward extent at 135°E is south of 31°N.

  • The LM is determined to be the ONLM if the longitude of its main body is east of 139°E and its southward extent at 140.5°E is south of 33°N.

  • The NNLM is determined to be the ONLM if its southward extent at 140.5°E is south of 33°N.

Appendix 3: Definition of the net Kuroshio transport

In considering the relation between the Kuroshio path and transport, the Kuroshio in an unperturbed state is treated as the basic state. Thus, the net or throughflow transport past the southern coast of Japan, excluding the contribution by eddies, is dynamically relevant. Here we explain the definition of the throughflow transport for the simulation, the linear model, and the repeat hydrographic observation along 137°E. Admittedly the definitions are simplistic, but the results compare well with each other.

3.1 Simulation

Ideally, the net Kuroshio transport should be defined at the saddle point along the stream function ridge of the subtropical gyre south of Japan (Fig. 19b). But a well-defined ridge structure is often deformed by mesoscale eddies. For the simulated fields, the throughflow transport south of Japan is obtained by the following procedure.

  1. 1.

    Evaluate the maximum of the barotropic stream function between the southern coast of Japan and 25°N along each meridional lattice point of the model from 135° to 141°E (values are plotted in Fig. 19a, and locations are on the blue line in Fig. 19b).

  2. 2.

    The throughflow transport is defined as the smallest of these values (circle in Fig. 19a).

Fig. 19
figure 19

Schematic that explains the definitions of throughflow transport used in this study. a The maximum of the barotropic stream function for 1996 (shades in b) between the southern coast of Japan and 25°N as a function of longitude. The circle at the lowest value of these maxima defines the throughflow transport for the simulation. b The barotropic stream function for 1996 of the simulation (shades) and the bathymetry (purple dashed contours; depth in km). The heavy blue line connects the locations of the maximum barotropic stream function between the southern coast of Japan and 25°N along each meridional grid line. Their values are plotted in a. The red line represents the meridian of 137°E along which repeat hydrographic observations are conducted. The green line from 26° to 32°N along 140°E represents the line along which the stream function from the linear model is averaged to assess the throughflow transport. c The barotropic stream function along 137°E (red line of b). The symbol X defines the throughflow transport for the hydrographic observation along 137°E. See "Appendix 3" for details

Note that the zero of the simulated stream function is placed on the main island of Japan.

3.2 Linear baroclinic vorticity model

The throughflow transport south of Japan for the linear baroclinic vorticity model ("Appendix 1") is defined as the mean value of the stream function between 26° and 32°N along 140°E (green line in Fig. 19b). The zero of this stream function is placed at the western boundary of the American continent. As shown in Fig. 4c, the maximum stream function in the subtropical gyre of the linear model appears around 26°–32°N south of Japan. The 140°E meridian is on the Izu Ridge and is the eastern entrance for the southwest-flowing recirculation gyre that eventually leaves the region south of Japan as the Kuroshio and the relatively small Tsushima warm current (∼2.5 Sv).

3.3 Repeat hydrographic section along 137°E

To obtain the throughflow transport across the vertical section along the 137°E meridian, we had to exclude its local enhancement by the anticyclonic eddy off Shikoku. The hydrographic observation analysis team of JMA uses the following procedure (T. Nakano, personal communication).

  1. 1.

    The eastward transport stream function is calculated by integrating the vertically integrated geostrophic eastward transport relative to a depth of 1,500 m southward from the southern coast of Japan.

  2. 2.

    Searching southward along the section from the southern coast of Japan, the net Kuroshio transport is obtained as the first minimum value after the maximum stream function is found (the symbol X in Fig. 19c).

3.4 Comparison

Figure 5b–d show time series. The simulated fields show considerable intra-annual variability reflecting the nonlinear effect, but there is also clear long-term variability that is largely tracked by the linear model (Fig. 5b). The correlation coefficient for the annual mean time series of the simulation and the linear model shown in Fig. 5c is 0.575 for 1948–2007. The effective degree of freedom based on the formula for red-noise processes (Bartlett 1935; Bretherton et al 1999) is 18 for this correlation, and the correlation coefficient at the 1 % significance level for 18 degrees of freedom is 0.561.

Figure 5d compares the method of “Simulation” section applied to the simulation (green line), the step 2 of “Repeat hydrographic section along 137°E” section applied to the simulation (magenta line), and the whole procedures of “Repeat hydrographic section along 137°E” section applied to the hydrographic observations (137°E) (blue line). The bias between the simulation and the observations arises from the fact that the observations use a level of no motion at 1,500 m. The simulation is expected to be larger by the amount of the missing part of the barotropic component in the observations. By definition, the method of “Simulation” section would yield a larger transport than applying the step 2 of “Repeat hydrographic section along 137°E” section to the simulation. The correlation coefficient for the 5-year running mean between and the simulation (green) and the observations (light blue) is 0.776 for 1969–2005. That for the 5-year running mean between the observations (light blue) and the result of applying the step 2 of “Repeat hydrographic section along 137°E” section to the simulation (magenta) is 0.859 for 1969–2005. The effective degree of freedom for the 5-year running mean data is determined here to be 8 (∼(2007–1967)/5). The correlation coefficient at the 1 % significance level for 8 degree of freedom is 0.765.

Appendix 4: Eddy identification and tracking procedure

Eddies are simply connected regions with large and positive values of a quantity (e.g., Isern-Fontanet et al. 2003)

$$ Q = -\left(\frac{\partial u} {\partial x} \right)^2 -\left(\frac{\partial v} {\partial x} \right) \left(\frac{\partial u} {\partial y} \right), $$
(6)

which measures the relative contribution of deformation and vorticity for an almost divergence free two-dimensional turbulent flow. This quantity corresponds to the Okubo-Weiss parameter (Okubo 1970; Weiss 1991) except for a factor of 4 and a global change of sign. Cyclonic eddies are identified as simply connected regions where Q > 2.0 × 10−12s−2 and vorticity is positive.

Using a series of 5-day mean outputs, eddies are identified on the basis of the horizontal velocity fields at 100 m depth. The center of an eddy is defined as the SSH maximum or minimum. Eddies are tracked in the following manner: If an eddy with the same sign of vorticity is found within 1.5° of the eddy’s center in the next time step, it is regarded as the same eddy. If two eddies from the previous time step reach the same eddy, tracking continues for the eddy with the longer life time.

Rights and permissions

Reprints and permissions

About this article

Cite this article

Tsujino, H., Nishikawa, S., Sakamoto, K. et al. Effects of large-scale wind on the Kuroshio path south of Japan in a 60-year historical OGCM simulation. Clim Dyn 41, 2287–2318 (2013). https://doi.org/10.1007/s00382-012-1641-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00382-012-1641-4

Keywords

Navigation