1. Introduction
The Kuroshio is a western boundary current of the North Pacific subtropical gyre, which advects warm water from the tropics and flows northeastward along the continental slope of the East China Sea and southern coast of Japan. It veers off the Japanese coast at around 35°N, forming the eastward jet known as the Kuroshio Extension (KE). The warm Kuroshio/KE releases large amounts of heat and moisture (Bond and Cronin 2008; Kubota et al. 2008) and has a sharp SST front along its poleward flank. Recent high-resolution satellite measurements and numerical simulations have revealed significant impacts of the Kuroshio/KE on the overlying marine atmospheric boundary layer (MABL), such as surface winds and surface air temperature (SAT), through heat and moisture exchanges (Xie et al. 2002; Nonaka and Xie 2003; Tokinaga et al. 2006, 2009; Small et al. 2008; Kelly et al. 2010; Minobe et al. 2010; Xu et al. 2011; Masunaga et al. 2016, 2020; Sugimoto et al. 2017). Responses to the Kuroshio/KE are also detectable in the free atmosphere, precipitation (Miyama et al. 2012; Sasaki et al. 2012; Sasaki and Yamada 2018; Xu et al. 2018), synoptic weather disturbances (Tanimoto et al. 2011; Nakamura et al. 2012; Hayasaki et al. 2013; Kuwano-Yoshida and Minobe 2017; Masunaga et al. 2020), and large-scale atmospheric circulation across the North Pacific (Nakamura et al. 2004; Ma et al. 2015, 2017; Qiu et al. 2014, 2020). Most of these previous studies have investigated the atmospheric responses to the Kuroshio/KE in autumn, winter, and spring, when the SST front becomes stronger and the heat and moisture exchanges become more active. In contrast, there are few studies on the effects of the Kuroshio/KE on the overlying atmospheric field in summer because of the weak SST front and small heat/moisture exchanges.
The Kuroshio south of Honshu, Japan, has remarkable bimodal features (Fig. 1a; Taft 1972; Kawabe 1985); a large meander (LM) path, detouring (i.e., meandering) offshore south of Japan, and a non-LM path, flowing along the southern coast of Japan. Both paths are relatively stable and maintained for several years to a decade once formed (Kawabe 1987). These features are not found in the other western boundary currents, such as the Gulf Stream. In the LM period, a developed cyclonic eddy is present in the onshore region between the Kuroshio and the southern coast of Tokai district (Fig. 1b). It has been recognized that the cool water pool is broadly distributed because of the upwelling and shoaling thermocline inside the cyclonic eddy. The LM-induced cool water pool can affect the overlying atmospheric field, including reduced surface winds in winter (Xu et al. 2010) and decreased precipitation in winter (Xu et al. 2010) and throughout the year (Murazaki et al. 2015). It also induces a significant southward shift of extratropical cyclone tracks in winter (Nakamura et al. 2012; Hayasaki et al. 2013). The cool water pool off Tokai district was considered the prominent feature in the LM path, and numerous air–sea-coupled studies have been conducted for the cold season (i.e., mainly in winter). A recent study by Sugimoto et al. (2020) showed marked coastal warming off Kanto–Tokai district during LM periods (Fig. 1c), which is attributable to the westward Kuroshio bifurcation that occurred during the LM periods, based on high-resolution satellite-derived data. The detected warming was in sharp contrast to the previous recognition that a cool water pool is distributed broadly in the region between the Kuroshio and the southern coast of Tokai district. Based on statistical analyses of weather station data, Kanto–Tokai district becomes warmer than usual in summer during LM periods, which indicates that LM-induced coastal warming could have an influence on the summertime climate in Japan via coastal air–sea interaction. However, the physical processes responsible for this have not yet been clarified.
It has long been recognized that the summer climate in Japan is largely affected by a combination of the North Pacific high, Bonin high (Ogasawara high), Tibetan high, and Okhotsk high. These highs are closely related to the large-scale atmospheric circulation fluctuations associated with four atmospheric teleconnection patterns: the Europe–Japan patterns (EJ1 and EJ2), which prevail over northern Eurasia and are linked to variations of the Okhotsk high (Wakabayashi and Kawamura 2004); the West Asia–Japan (WJ) pattern, which is characterized by a stationary wave train pattern along the upper-level subtropical jet from West Asia to the central North Pacific (Wakabayashi and Kawamura 2004) that is similar to the Silk Road pattern identified by Enomoto et al. (2003); and the Pacific–Japan (PJ) pattern, which is a stationary wave train pattern generated by tropical convective activity in the vicinity of the Philippine Sea (e.g., Kurihara and Tsuyuki 1987; Nitta 1987; Kosaka and Nakamura 2010). It has been reported that the WJ pattern influences the surface air temperature in the western part of Japan, and the PJ, EJ1, and EJ2 patterns affect the SAT in the northern part of Japan (Tachibana et al. 2004; Wakabayashi and Kawamura 2004; Kubota et al. 2016). With recent progress in regional atmospheric modeling, it has been shown that the SST around Japan can exert an influence on summer climate (e.g., Takahashi et al. 2015); the northward movement of the SST front to the east of Japan triggers a northward shift of the tropospheric jet (Nakamura and Miyama 2014; Matsumura et al. 2016), resulting in anomalous warming in northern Japan (Nakamura and Yamane 2010; Matsumura et al. 2016); the LM-induced cool water pool off Tokai district induces lower SAT and increased precipitation on the Pacific coast of Japan (Murazaki et al. 2015). Most studies have focused on the effects of the open ocean, large-scale SST (>500 km) on the climate of Japan, and our understanding of the smaller-scale SST impacts, such as coastal warming, is still lacking.
It is expected that the coastal warming off Kanto–Tokai attributable to the LM path would exert a significant impact on summer climate, because the southerly wind flow along the periphery of the western Pacific subtropical high (Fig. 2a) is predominant over Japan in summer. In addition, the summertime SST varies considerably off Kanto–Tokai district with coastal warming (Fig. 2b). The purpose of this study is to reveal summertime local atmospheric responses to coastal warming off Kanto–Tokai district, and quantitatively assess its remote effects on the summer climate of Japan via low-level atmospheric processes. We conduct regional atmospheric model experiments using the recently developed SST product on a 1/100° (longitude) × 1/100° (latitude) grid from the NASA Jet Propulsion Laboratory, which adequately captures the coastal warming. In the summer of 2017, the Kuroshio took the LM path for the first time since the summer of 2005, and this event remains ongoing as of January 2021, at the time of writing of this manuscript. Our study is timely and will be helpful for predicting the effects of the ongoing LM event on the summer climate of Japan.
2. Data and model
a. Observations and reanalysis data
We use the daily NASA Jet Propulsion Laboratory multiscale, ultra-high-resolution SST (MURSST) product (Chin et al. 2013) on a 1/100° (longitude) × 1/100° (latitude) grid. This product incorporates SSTs from eight satellites, using both infrared and passive microwave retrievals, along with in situ data. It has been used previously for research on coastal upwelling (Vazquez-Cuervo et al. 2013; Gentemann et al. 2017). Data from this product are available from January 2003 onward.
We use the monthly SAT data at weather stations and the Automated Meteorological Data Acquisition System (AMeDAS) stations operated by the Japan Meteorological Agency (JMA). We also use four atmospheric teleconnection pattern indices, calculated from the JMA Japanese 55-year Reanalysis (JRA-55) data (Kobayashi et al. 2015), based on the definitions of Wakabayashi and Kawamura (2004): PJ, WP, EJ1, and EJ2. The Niño-3 index (SST anomalies averaged for 5°S–5°N, 150°–90°W), is also used, which is regarded as an indicator of El Niño events.
We examine 18 summers (June–August) from 2003–20, for which the above datasets are available.
b. Regional atmospheric model experiment
The regional atmospheric model used in this study is the JMA/Meteorological Research Institute (JMA/MRI) nonhydrostatic model (NHM; Saito et al. 2006, 2007). The NHM is designed to serve both weather forecasting and atmospheric research needs, and has been used in a variety of regional climate studies. The model domain covers the Kuroshio south of Japan from 28° to 40°N and from 130° to 145°E in the Lambert projection (Fig. 3a), with 5-km horizontal grid spacing. The model has 51 vertical levels with realistic topography; 19 levels are placed below 2 km in height to finely resolve the MABL. We used the Mellor–Yamada–Nakanishi–Niino level-3 planetary boundary layer scheme (Nakanishi and Niino 2004), the Beljaars–Holtslag flux and bulk coefficient scheme (Beljaars and Holtslag 1991), and the Kain–Fritsch convection parameterization scheme (Kain and Fritsch 1993; Kain 2004) for convective processes.
The coastal warming off Kanto–Tokai district ranges from 1.5° to 3.5°C during the latest LM event (Fig. 2c). In our experiments, we focus on the month of July, when the baiu rainband is located in northern Japan away from the coastal area off Kanto–Tokai district, as compared with June when the rainband is found around southern part of Japan and August when more typhoons approach Japan. To detect the effects of the coastal warming off Kanto–Tokai district on the overlying atmosphere and climate of Japan, we conduct two sets of experiments by imposing different SST boundary conditions. The first is the control (CTRL) run, in which we prescribe daily SSTs from MURSST for 17 June to 31 July 2004, when the Kuroshio takes the LM path, and in a near-neutral phase of natural variability, such as El Niño and the PJ pattern, that affect the summertime atmospheric conditions around Japan. The second experiment (COLD run) is conducted by subtracting the idealized SST anomaly (Fig. 3b), representing the coastal warming, from the daily SST field of the CTRL run. We used the idealized SST because a simple difference in SST between the LM and non-LM periods includes effects that are not related to the Kuroshio LM path, such as eddies. The idealized SST anomaly has a maximum amplitude of about 3°C, which falls within +1 SD of the mean value (Fig. 2c). This experiment design differs from the traditional method of using SST data and spatially and/or temporally smoothed data as the boundary condition (e.g., Takahashi et al. 2015; Matsumura et al. 2016). In addition, SSTs used as a boundary condition in previous studies do not resolve coastal warming. Our experiment is designed to reveal only the atmospheric response to coastal warming off Kanto–Tokai district.
For the initial and boundary atmospheric conditions, we use the recently released fifth generation of atmospheric reanalysis of the global climate of ECMWF (ERA5; Copernicus Climate Change Service 2017), which is on a 1/4° (longitude) × 1/4° (latitude) grid. The CTRL and COLD experiments consist of seven NHM simulations, in which the initial and boundary conditions were calculated based on the method of Inatsu and Terakura (2012). The period between 17 June and the end of the month is a spinup phase, and then July is analyzed. All results are shown as an average over July and an average over the ensemble. It is worth emphasizing that in this regional modeling approach, the initial and lateral boundary conditions are identical in the two experiments.
3. Local atmospheric response to coastal warming off Kanto–Tokai district
The CTRL run from NHM (Fig. 4b) successfully captures the large-scale features of the summer SAT field in the ERA5 (Fig. 4a). The SAT fields show a gradual westward increase south of 35°N and a sharp meridional front associated with the KE along 35°N east of Japan. Low SATs are detected north of 35°N east of Japan, which reflects the southward cold-water intrusion associated with the Oyashio. Over the islands of Japan, the CTRL run shows higher and lower temperature in some regions, compared with the ERA5, because of the spatially high-resolution topography. The surface winds show a striking correspondence between ERA5 and the CTRL run; southwesterly winds are dominant south of Japan and blow strongly into Kanto district. These provide confidence that the NHM simulation is capable of successfully reproducing climate features over the Kuroshio south of Honshu, Japan, and Pacific coast cities, such as in Kanto district, in July. However, our simulation may depend on the spatial resolution and parameterization.
a. Air temperature and water vapor response
We explore the local atmospheric responses over the coastal warming off Kanto–Tokai district attributable to the LM path. A comparison between the CTRL and COLD runs shows a local SAT increase off Kanto–Tokai district (Fig. 5a), with the values increasing by 3°C in the CTRL run, which is almost consistent with the SST increase in terms of both location and amplitude (Fig. 3b). We investigate the upward influence of the coastal warming by examining the difference in virtual potential temperature along 34°N, where a marked difference in SAT can be detected, between the two runs (Fig. 5b). Significant warming extends not only into the near surface but also throughout the MABL, reaching the 950-hPa level in the CTRL run.
Numerous studies have noted that the summertime SST anomalies in the western North Pacific are formed by downward solar heating (e.g., Wu and Kinter 2010). In contrast to the basin-scale studies, our recent work (Sugimoto et al. 2020) showed that the SST increase off Kanto–Tokai district is not related to downward solar heating, but is associated with advection of warm Kuroshio water by meander and the resulting enhanced upward heat flux in summer. The upward heat flux occurs predominantly by latent heat flux and, to a lesser extent, by sensible heat flux. Therefore, it is expected that the increase in upward heat fluxes can induce warming of the MABL. Our experiments reveal a collocation of increased SATs and enhanced upward heat release off Kanto–Tokai district (Figs. 5a and 6a,b). The heat flux increase attributable to the coastal warming is about 100 W m−2, and the latent heat flux is about 10 times larger than the sensible heat flux. We diagnostically investigate the diabatic heating, following Yanai et al. (1973) and Yanai and Tomita (1998), to detect the effects of the oceanic heat release on the overlying atmosphere. There is a strong diabatic heating rate over the coastal warming area (Fig. 6c), the effect of which extends to the top of the MABL (Fig. 6d). It is evident that the coastal warming off Kanto–Tokai district heats the MABL both directly and locally.
The large latent heat flux (Fig. 6a)—which is enhanced evaporation from the ocean—over the coastal warming area results in an increase in water vapor in the MABL (Figs. 7a,b). It is expected that the large amount of water vapor affects the cloud distribution. Figure 7d shows the difference in cloud amount along 34°N between the two runs, indicating a well-defined seesaw pattern in the vertical direction, with increased cloudiness between the 970- and 920-hPa levels and reduced cloudiness at the near surface. In the upper part, the cloud increases (Fig. 7c) are collocated with the area of enhanced evaporation (Fig. 6a), except for a slight displacement of the cloud maximum downstream of the evaporation, which is consistent with advection by the prevailing southwesterly winds. The lower part implies a fog formation due to the lack of coastal warming during the COLD run (Fig. 7e).
b. Surface wind response
The recent observational study of Schneider (2020) indicated that the ocean mesoscale SST has a significant impact on near-surface winds in the midlatitude ocean. It is expected that the coastal warming off Kanto–Tokai district attributable to the LM path also has an impact on winds. A difference in surface winds between the CTRL and COLD runs (Fig. 8a) clearly shows westerly wind anomalies over the coastal warming area, revealing wind acceleration due to the consistency in wind direction with the mean state (Fig. 4b). In addition, westerly wind anomalies are detected within the MABL and, in contrast, easterly wind anomalies are found above the MABL height (Fig. 8b).
It is important to explore what processes contribute to the surface wind responses to the coastal warming. Two processes have been proposed as effects of SST, at SSTs below the convective threshold (i.e., 26°–27°C): 1) the pressure adjustment process (Lindzen and Nigam 1987; Minobe et al. 2008), whereby a depression in sea level pressure (SLP) around a positive SST anomaly forms a pressure gradient, resulting in surface wind changes; and 2) the vertical mixing process (Wallace et al. 1989; Tanimoto et al. 2011), which produces surface wind changes due to a decrease in static stability above positive SST anomalies that enhance vertical mixing, and promote momentum transport from above into the lower boundary layer or through equilibrium changes in boundary layer height (Samelson et al. 2006). These processes have been mainly investigated over oceanic western boundary currents with sharp SST fronts, such as the KE and Gulf Stream.
Our results reveal that the westerly wind anomalies are dominated by the vertical mixing term (Fig. 9d), which reflects the downward transfer of westerly component momentum owing to the monsoonal westerly component aloft in summer. This results in the acceleration of surface westerly winds. The pressure gradient term is toward the north over the coastal warming area (Fig. 9a), indicating the wind response to the negative SLP anomalies attributable to the coastal warming. In contrast, the Coriolis term is toward the south (Fig. 9b) over the coastal warming area, and acts to cancel the pressure gradient–driven flow because of the identical amplitude compared with the pressure gradient term (Fig. 9c).
4. Kanto district warming linked to the Kuroshio LM path via a local greenhouse effect
A close look at Fig. 10a reveals significant warming on the Pacific coast of Kanto and Tokai districts, which is attributable to the coastal warming off Kanto–Tokai district. Here we investigate the remote impact of coastal warming on summer climate in Kanto district, where the megacity Tokyo is located and >40 million people live.
In Kanto district (KD; a land area surrounded by a blue line in Fig. 10a), the SAT in the CTRL run is 0.61°C warmer than in the COLD run. The SAT increase could be attributable to two processes: 1) horizontal temperature advection in the lower troposphere and 2) net radiation heating at the surface. Figure 10b displays the difference in horizontal temperature advection at the surface between the CTRL and COLD runs. Positive anomalies can be observed in the Pacific coastal regions, including Kanto district, but are not significant. We checked that the results were similar at other heights. In contrast, the net radiation at the surface [i.e., the sum of net shortwave radiation flux and net longwave radiation (LWR) flux] shows significant positive anomalies in the Pacific coastal regions, including Kanto district (Fig. 10c), which are collocated with the SAT pattern in Fig. 10a. By comparing the four components of net radiation (Fig. 11), it is found that the downward LWR (Fig. 11a) has the largest amplitude, and which also has a spatial pattern similar to the net radiation, with an amplitude that increases toward the Pacific coast. The downward LWR is responsible for the determination of net radiation. The increase in downward LWR in the KD is 3.3 W m−2. Previous studies have noted that solar heating has an essential role in the summer SATs over Japan (e.g., Yasunaka and Hanawa 2006), but the solar radiation response to coastal warming is incoherent in our simulation (Fig. 11c), which is consistent with the incoherent cloud responses over Japan to coastal warming (Fig. 7f).
Changes in downward LWR at the surface largely reflect cloudiness and water vapor in the atmosphere. The changes in cloudiness were incoherent over Kanto district. Accordingly, we focus on the water vapor [i.e., total precipitable water (TPW)]. A comparison between the CTRL and COLD runs shows significant increases in TPW in Kanto–Tokai district (Fig. 12a), which are more prominent on the Pacific coast side. The increase in TPW in KD is 2.3 mm. These spatial features are very similar to that in the downward LWR difference in Fig. 11a. Qualitatively, it appears that the increase in TPW induces the increase in downward LWR at the surface, resulting in the increase in SATs in Kanto district.
We attempt to quantitatively evaluate the influence of TPW on the downward LWR and SAT based on a diagnosed approach. We adopt the approach for clear sky proposed by Allan et al. (2004) because our simulation showed an incoherent cloudiness response to coastal warming. We confirmed that identical results to Fig. 11a were obtained from the clear-sky LWR in our simulation (not shown). In the diagnostic calculation, July-mean values averaged within the KD in the CTRL run are set as a basic state: SAT = 299 K, TPW = 47.7 mm, and downward LWR at surface = 402.2 W m−2. Given the TPW difference between the two runs (2.3 mm), the diagnosed increase in downward LWR is about 2.3 W m−2, and these conditions result in an increase in SAT of about 0.43 K. The diagnosed LWR and SAT responses to the TPW increase are about 60% and 70% of the simulated results (3.3 W m−2 and 0.61 K), respectively. It is apparent that the TPW is primarily responsible for the SATs in Kanto district.
These results indicate that the combination of low-level wind circulation changes and increase in water vapor attributable to the coastal warming off Kanto–Tokai district leads to water vapor convergence over Kanto district. This causes the increase in downward LWR at the surface and the surface air warming in response to a local greenhouse effect.
5. Observational evidence of hot summers in Kanto district attributable to the Kuroshio LM path
a. Large SAT increase in Kanto district attributable to the Kuroshio LM path
Our simulation revealed that the coastal warming off Kanto–Tokai district attributable to the Kuroshio LM path leads to a SAT increase in KD of ~0.6°C. To evaluate the simulated warming, we compare it with the observed anomaly during the present LM event. Figure 14 displays summertime SAT anomalies for the latest LM event. The anomalies are different from mean values during the non-LM period of 2006–16, as in imposed SST in our simulation. During the latest LM event, the summertime climate has been affected by various phenomena. For example, Japan experienced an unprecedented heat wave with record-high temperatures in July 2018 (Imada et al. 2019), and record-heavy rainfall and record-low sunshine durations due to the active mei-yu–baiu front in July 2020, resulting in low temperatures (JMA 2020). Nevertheless, the SAT fields during the latest LM path period show large positive anomalies around Kanto district of ~0.4°C in KD. This is the same order of amplitude as the simulated LM-induced warming. This indicates that the Kuroshio path can significantly impact the summer climate in Kanto district.
b. Coastal warming versus teleconnection and climate patterns
We revealed the significant impacts of coastal warming off Kanto–Tokai district on summer SATs in Kanto district through SST sensitivity experiments. To verify the model result, we examine its relationship with an observational dataset. Figure 15a displays a map of correlation coefficients between SST off Kanto–Tokai district (green dot in the lower panel in Fig. 15a) and SATs from weather station/AMeDAS data in summer. Significant relationships are concentrated predominantly in the Pacific coast of Kanto and Tokai districts, which is consistent with our simulated results (Fig. 10a).
Numerous studies have noted that various atmospheric teleconnection patterns and El Niño events affect the summertime SATs over Japan. We attempt to identify the significance of the oceanic impact on the summer climate of Japan by comparisons with teleconnection patterns and El Niño. The PJ pattern is significantly correlated with the SATs in eastern Japan (Fig. 15b), and El Niño shows a significant relationship in western Japan (Fig. 15f), but these have little effect on the climate in Kanto and Tokai districts. For the WJ, EJ1, and EJ2 patterns, there are hardly any significant signals during this period (Figs. 15c–e). These observational results highlight us the uniqueness of the Kuroshio path changes on the summertime climate variability on the Pacific coast of Kanto and Tokai districts, which supports our simulation results.
6. Discussion
Our experiments showed that the change in downward LWR at the surface associated with the water vapor change is the dominant influence on the summertime SATs in Kanto district. It is known that the LWR at the surface has diurnal variations, such as upward LWR (radiative cooling) during nighttime. We examine the diurnal sensitivity to the LM-induced coastal warming off Kanto–Tokai district in our experiments. Figure 16 displays the diurnal difference in SATs between the two runs in Kanto district, as a function of Japan standard time (UTC + 9 h). A diurnal variation is evident, with a high sensitivity of 0.7°C from midnight to early morning and low sensitivity during the day (0.4°C). This indicates that the daily minimum SAT tends to increase during the LM path periods, which would be caused by the reduced radiative cooling associated with the increased water vapor. The temporal resolution of the SST data used in our experiments is daily, and the effects of diurnal SST variation on the regional climate remains unclear. Some studies denoted that the diurnal SST variation in some coastal regions has a large amplitude [see the review of Kawai and Wada (2007)]. For example, the diurnal SST amplitude is as large as 5.5°C in Tokyo Bay in summer (Oda and Kanda 2009). Ocean assimilation models have improved and progressed dramatically in recent years, and have a temporal resolution of 30 min and 1 h with a spatial resolution of 1–2 km. These new assimilated SST data should help us in the future to quantitatively assess the influence of LM-induced coastal warming on the regional climate.
We have shown that the LM-induced coastal warming off Kanto–Tokai district exerted significant remote impacts on the summer climate in Kanto district via the southerly winds that flow along the periphery of the western Pacific subtropical high and are predominant over Japan in summer. Does the coastal warming have an influence on the climate in winter when the northerly winds are dominant due to the west–high, east–low pressure pattern (Fig. 17a)? To get hints to this, we examine the relationship between SATs and coastal warming in winter. Figure 17b shows no significant signals over Japan, including Kanto district. This result emphasizes that the Kuroshio’s impacts on the SATs over Japan become apparent only in summer.
7. Summary and concluding remarks
The westward Kuroshio bifurcation induces coastal warming off Kanto–Tokai district during LM periods, with SST anomalies of 1.5°–3.5°C during the latest LM path period of 2017–20, compared with the non-LM path period of 2006–16. We conducted an atmospheric response sensitivity experiment to the idealized SST with a maximum amplitude of 3°C that represents the coastal warming. Our regional atmospheric model experiments revealed that the summertime coastal warming off Kanto–Tokai district attributable to the Kuroshio LM path heats the MABL both directly and locally through a large upward heat release, which then increases the water vapor in the MABL through enhanced evaporation from the ocean. A momentum budget analysis showed that the coastal warming leads to the acceleration of westerly winds in the near-surface atmosphere via a vertical mixing process, owing to the presence of monsoonal westerly winds aloft in summer, and to the development of easterly/southerly winds blowing into Kanto district due to a combination of a pressure adjustment process and Coriolis force in the east of the warming area. Furthermore, our simulations and supplementary observational studies revealed the significant remote impacts of coastal warming on summertime SATs in Kanto district. The Kanto district was 0.6°C warmer due to coastal warming, which is similar to the observed SAT anomalies during the latest LM period (+0.4°C than the non-LM periods of 2006–16), even though the observations include effects other than the LM. This indicates that the Kuroshio path state has a significant impact on the summer climate in Kanto district. A SAT increase results from a combination of the low-level wind circulation changes and increase in water vapor attributable to coastal warming off Kanto–Tokai district. This leads to the water vapor convergence over Kanto district, resulting in an increase in downward LWR at the surface and then surface air warming due to a local greenhouse effect. Atmospheric teleconnection patterns and El Niño events have no significant impacts on the SATs in Kanto district, which also highlights the uniqueness of the Kuroshio path state on the summertime climate variability in Kanto district.
Kanto district, including Tokyo, is one of the 37 megacities in the world and has 40 million inhabitants. Its summer climate is characterized by high temperature and humidity that lead to discomfort for its inhabitants. Our experiments showed that the climate in Kanto district becomes hotter and more humid because of the Kuroshio LM path. We attempt to provide an additional perspective on the impacts of Kuroshio on the summer climate in Japan in terms of discomfort information. Here, as an indicator of discomfort, we use a modified version for Japanese climate (Kinouchi 2001; Table 1) of Thom’s index (Thom 1957, 1959), based on the basic environmental parameters SAT and relative humidity. The mean state in the CTRL run yields remarkably high discomfort values in Kanto district (Fig. 18a). To examine the influence of the Kuroshio on the discomfort in further detail, we count the days that the daily maximum discomfort index exceeds 80 (i.e., most inhabitants feel discomfort). As expected, the difference between the two runs is remarkable in Kanto district, and is 13.1 days in the CTRL run, which is 160% greater than that in the COLD run (8.1 days; Fig. 18b). The Kuroshio path state is responsible for the climatic comfort of living in Tokyo, Japan.
Classification of human comfort during summer according to discomfort index (DI) values.
In the summer of 2017, the Kuroshio took the LM path, and this event is ongoing as of January 2021. The establishment of a general and regional framework that describes how LM-induced coastal warming drives the atmosphere could greatly improve weather and seasonal forecasting. Our approach is also applicable in other coastal cities worldwide.
Acknowledgments
The authors thank Eitarou Oka, Kaoru Ichikawa, Shin Fukui, Ryusuke Masunaga, and Akira Yoshida for valuable comments. Constructive comments made by two anonymous reviewers have significantly improved an early version of the manuscript. The ERA-5 data used in this study were accessed from https://cds.climate.copernicus.eu, the JMA weather stations and AMeDAS stations from http://www.jma.go.jp/jma/indexe.html, and the NASA Jet Propulsion Laboratory MURSST data from https://opendap.jpl.nasa.gov. Our numerical experiments were conducted with the JMA–NHM developed by the Meteorological Research Institute and Numerical Prediction Division of the Japan Meteorological Agency. Part of the experimental results was obtained using supercomputing resources at the Cyberscience Center, Tohoku University. This study was conducted in JFY2019 when SS was staying at the University of Hawaii under the Tohoku University 2019 Leading Young Researcher Overseas Visit Program, and SS thanks Kelvin Richards, Toshiyuki Hayase, Ryota Hino, and Toshio Suga for their help. SS was supported by JSPS Grant 18K03737 and MEXT Grant 19H05704, and Japan Fisheries Research and Education Agency. BQ was supported by NSF Grant 2019312. NS was supported by the U.S. National Aeronautics and Space Administration Grant 80NSSC19K0058 and by the JAMSTEC IPRC Collaborative Research (JICore).
REFERENCES
Allan, R. P., M. A. Ringer, J. A. Pamment, and A. Slingo, 2004: Simulation of the Earth’s radiation budget by the European Centre for Medium-Range Weather Forecasts 40-year reanalysis (ERA40). J. Geophys. Res., 109, D18107, https://doi.org/10.1029/2004JD004816.
Beljaars, A., and A. Holtslag, 1991: Flux parameterization over land surfaces for atmospheric models. J. Appl. Meteor., 30, 327–341, https://doi.org/10.1175/1520-0450(1991)030<0327:FPOLSF>2.0.CO;2.
Bond, N. A., and M. F. Cronin, 2008: Regional weather patterns during anomalous air–sea fluxes at the Kuroshio Extension Observatory (KEO). J. Climate, 21, 1680–1697, https://doi.org/10.1175/2007JCLI1797.1.
Chin, T. M., J. Vazquez, and E. Armstrong, 2013: A multi-scale, high-resolution analysis of global sea surface temperature. Algorithm Theoretical Basis Document, version 1, 13.
Copernicus Climate Change Service (C3S), 2017: ERA5: Fifth generation of ECMWF atmospheric reanalyses of the global climate, Copernicus Climate Change Service Climate Data Store (CDS), last accessed 25 March 2019, https://cds.climate.copernicus.eu/#!/search?text=ERA5&type=dataset.
Enomoto, T., B. J. Hoskins, and Y. Matsuda, 2003: The formation mechanism of the Bonin high in August. Quart. J. Roy. Meteor. Soc., 129, 157–178, https://doi.org/10.1256/qj.01.211.
Gentemann, C. L., M. R. Fewings, and M. García-Reyes, 2017: Satellite sea surface temperatures along the West Coast of the United States during the 2014–2016 northeast Pacific marine heat wave. Geophys. Res. Lett., 44, 312–319, https://doi.org/10.1002/2016GL071039.
Hayasaki, M., R. Kawamura, M. Mori, and M. Watanabe, 2013: Response of extratropical cyclone activity to the Kuroshio large meander in northern winter. Geophys. Res. Lett., 40, 2851–2855, https://doi.org/10.1002/grl.50546.
Imada, Y., M. Watanabe, H. Kawase, H. Shiogama, and M. Arai, 2019: The July 2018 high temperature event in Japan could not have happened without human-induced global warming. SOLA, 15A, 8–12, https://doi.org/10.2151/sola.15A-002.
Inatsu, M., and K. Terakura, 2012: Wintertime extratropical cyclone frequency around Japan. Climate Dyn., 38, 2307–2317, https://doi.org/10.1007/s00382-011-1152-8.
Ito, T., 2019: A study on water vapor transport to Tohoku region regarding the Kanto Tohoku heavy rainfall in September, 2015 (in Japanese). M.S. thesis, Department of Geophysics, Tohoku University, 65 pp.
JMA, 2020: Climate characteristics of record-heavy rain and record-low sunshine durations in Japan in July 2020. JMA, 16 September 2020, https://ds.data.jma.go.jp/tcc/tcc/news/press_20200916.pdf.
Kain, J., 2004: The Kain–Fritsch convective parameterization: An update. J. Appl. Meteor., 43, 170–181, https://doi.org/10.1175/1520-0450(2004)043<0170:TKCPAU>2.0.CO;2.
Kain, J., and J. Fritsch, 1993: Convective parameterization for meso-scale models: The Kain–Fritsch scheme. The Representation of Cumulus Convection in Numerical Models, Meteor. Monogr., Amer. Meteor. Soc., 46, 165–170.
Kawabe, M., 1985: Sea level variations at the Izu Islands and typical stable paths of the Kuroshio. J. Oceanogr. Soc. Japan, 41, 307–326, https://doi.org/10.1007/BF02109238.
Kawabe, M., 1987: Spectral properties of sea level and time scales of Kuroshio path variations. J. Oceanogr. Soc. Japan, 43, 111–123, https://doi.org/10.1007/BF02111887.
Kawai, Y., and A. Wada, 2007: Diurnal sea surface temperature variation and its impact on the atmosphere and ocean: A review. J. Oceanogr., 63, 721–744, https://doi.org/10.1007/s10872-007-0063-0.
Kelly, K. A., R. J. Small, R. M. Samelson, B. Qiu, T. M. Joyce, Y.-O. Kwon, and M. F. Cronin, 2010: Western boundary currents and frontal air–sea interaction: Gulf Stream and Kuroshio Extension. J. Climate, 23, 5644–5667, https://doi.org/10.1175/2010JCLI3346.1.
Kinouchi, T., 2001: A study on thermal indices for the outdoor environment (in Japanese). Tenki, 48, 661–671.
Kobayashi, S., and Coauthors, 2015: The JRA-55 reanalysis: General specifications and basic characteristics. J. Meteor. Soc. Japan, 93, 5–48, https://doi.org/10.2151/jmsj.2015-001.
Kosaka, Y., and H. Nakamura, 2010: Mechanisms of meridional teleconnection observed between a summer monsoon system and a subtropical anticyclone. Part I: The Pacific–Japan pattern. J. Climate, 23, 5085–5108, https://doi.org/10.1175/2010JCLI3413.1.
Kubota, H., Y. Kosaka, and S.-P. Xie, 2016: A 117-year long index of the Pacific–Japan pattern with application to interdecadal variability. Int. J. Climatol., 36, 1575–1589, https://doi.org/10.1002/joc.4441.
Kubota, M., N. Iwabe, M. F. Cronin, and H. Tomita, 2008: Surface heat fluxes from the NCEP/NCAR and NCEP/DOE reanalyses at the Kuroshio Extension Observatory buoy site. J. Geophys. Res., 113, C02009, https://doi.org/10.1029/2007JC004338.
Kurihara, K., and T. Tsuyuki, 1987: Development of the barotropic high around Japan and its association with Rossby wave-like propagations over the North Pacific: Analysis of August 1984. J. Meteor. Soc. Japan, 65, 237–246, https://doi.org/10.2151/jmsj1965.65.2_237.
Kuwano-Yoshida, A., and S. Minobe, 2017: Storm-track response to SST fronts in the northwestern Pacific region in an AGCM. J. Climate, 30, 1081–1102, https://doi.org/10.1175/JCLI-D-16-0331.1.
Lindzen, R. S., and R. S. Nigam, 1987: On the role of sea surface temperature gradients in forcing low-level winds and convergence in tropics. J. Atmos. Sci., 44, 2418–2436, https://doi.org/10.1175/1520-0469(1987)044<2418:OTROSS>2.0.CO;2.
Ma, X., P. Chang, R. Saravanan, and R. Montuoro, 2015: Distant influence of Kuroshio eddies on North Pacific weather patterns? Sci. Rep., 5, 17785, https://doi.org/10.1038/srep17785.
Ma, X., P. Chang, R. Saravanan, R. Montuoro, H. Nakamura, and D. Wu, 2017: Importance of resolving Kuroshio front and eddy influence in simulating the North Pacific storm track. J. Climate, 30, 1861–1880, https://doi.org/10.1175/JCLI-D-16-0154.1.
Masunaga, R., H. Nakamura, T. Miyasaka, K. Nishi, and B. Qiu, 2016: Interannual modulations of oceanic imprints on the wintertime atmospheric boundary layer under the changing dynamical regimes of the Kuroshio Extension. J. Climate, 29, 3273–3296, https://doi.org/10.1175/JCLI-D-15-0545.1.
Masunaga, R., H. Nakamura, B. Taguchi, and T. Miyasaka, 2020: Processes shaping the time-mean surface wind convergence patterns in winter around the Kuroshio Extension and Gulf Stream. J. Climate, 33, 3–25, https://doi.org/10.1175/JCLI-D-19-0097.1.
Matsumura, S., T. Horinouchi, S. Sugimoto, and T. Sato, 2016: Response of the baiu rainband to northwest Pacific SST anomalies and its impact on atmospheric circulation. J. Climate, 29, 3075–3093, https://doi.org/10.1175/JCLI-D-15-0691.1.
Mertz, F., M. I. Pujol, and Y. Faugére, 2018: Product user manual (CMEMS-SL812 PUM-008-032-051), version 4, 46 pp.
Minobe, S., A. Kuwano-Yoshida, N. Komori, S.-P. Xie, and R. J. Small, 2008: Influence of the Gulf Stream on the troposphere. Nature, 452, 206–209, https://doi.org/10.1038/nature06690.
Minobe, S., M. Miyashita, A. Kuwano-Yoshida, H. Tokinaga, and S.-P. Xie, 2010: Atmospheric response to the Gulf Stream: Seasonal variations. J. Climate, 23, 3699–3719, https://doi.org/10.1175/2010JCLI3359.1.
Miyama, T., M. Nonaka, H. Nakamura, and A. Kuwano-Yoshida, 2012: A striking early-summer event of a convective rainband persistent along the warm Kuroshio in the East China Sea. Tellus, 64A, 18962, https://doi.org/10.3402/tellusa.v64i0.18962.
Murazaki, K., H. Tsujino, T. Motoi, and K. Kurihara, 2015: Influence of the Kuroshio Large Meander on the climate around Japan based on a regional climate model. J. Meteor. Soc. Japan, 93, 161–179, https://doi.org/10.2151/jmsj.2015-009.
Nakamura, H., T. Sampe, Y. Tanimoto, and A. Shimpo, 2004: Observed associations among storm tracks, jet streams and midlatitude oceanic fronts. Earth’s Climate: The Ocean–Atmosphere Interaction, Geophys. Monogr., Vol. 147, Amer. Geophys. Union, 329–345.
Nakamura, H., A. Nishina, and S. Minobe, 2012: Response of storm tracks to bimodal Kuroshio path states south of Japan. J. Climate, 25, 7772–7779, https://doi.org/10.1175/JCLI-D-12-00326.1.
Nakamura, M., and S. Yamane, 2010: Dominant anomaly patterns in the near-surface baroclinicity and accompanying anomalies in the atmosphere and oceans. Part II: North Pacific basin. J. Climate, 23, 6445–6467, https://doi.org/10.1175/2010JCLI3017.1.
Nakamura, M., and T. Miyama, 2014: Impacts of the Oyashio temperature front on the regional climate. J. Climate, 27, 7861–7873, https://doi.org/10.1175/JCLI-D-13-00609.1.
Nakanishi, M., and H. Niino, 2004: An improved Mellor–Yamada level-3 model with condensation physics: Its design and verification. Bound.-Layer Meteor., 112 (1), 1–31, https://doi.org/10.1023/B:BOUN.0000020164.04146.98.
Nitta, T., 1987: Convective activities in the tropical western Pacific and their impact on the Northern Hemisphere summer circulation. J. Meteor. Soc. Japan, 65, 373–390, https://doi.org/10.2151/jmsj1965.65.3_373.
Nonaka, M., and S.-P. Xie, 2003: Covariations of sea surface temperature and wind over the Kuroshio and its extension: Evidence for ocean-to-atmosphere feedback. J. Climate, 16, 1404–1413, https://doi.org/10.1175/1520-0442(2003)16<1404:COSSTA>2.0.CO;2.
Oda, R., and M. Kanda, 2009: Observed sea surface temperature of Tokyo Bay and its impact on urban air temperature. J. Appl. Meteor. Climatol., 48, 2054–2068, https://doi.org/10.1175/2009JAMC2163.1.
Qiu, B., S. Chen, N. Schneider, and B. Taguchi, 2014: A coupled decadal prediction of the dynamic state of the Kuroshio Extension system. J. Climate, 27, 1751–1764, https://doi.org/10.1175/JCLI-D-13-00318.1.
Qiu, B., S. Chen, N. Schneider, E. Oka, and S. Sugimoto, 2020: On reset of the wind-forced decadal Kuroshio Extension variability in late 2017. J. Climate, 33, 10 813–10 828, https://doi.org/10.1175/JCLI-D-20-0237.1.
Saito, K., and Coauthors, 2006: The operational JMA non-hydrostatic mesoscale model. Mon. Wea. Rev., 134, 1266–1298, https://doi.org/10.1175/MWR3120.1.
Saito, K., J.-I. Ishida, K. Aranami, T. Hara, T. Segawa, M. Narita, and Y. Honda, 2007: Nonhydrostatic atmospheric models and operational development at JMA. J. Meteor. Soc. Japan, 85B, 271–304, https://doi.org/10.2151/jmsj.85B.271.
Samelson, R. M., E. D. Skyllingstad, D. B. Chelton, S. K. Esbensen, L. W. O’Neill, and N. Thum, 2006: On the coupling of wind stress and sea surface temperature. J. Climate, 19, 1557–1566, https://doi.org/10.1175/JCLI3682.1.
Sasaki, Y. N., and Y. Yamada, 2018: Atmospheric response to interannual variability of sea surface temperature front in the East China Sea in early summer. Climate Dyn., 51, 2509–2522, https://doi.org/10.1007/s00382-017-4025-y.
Sasaki, Y. N., S. Minobe, T. Asai, and M. Inatsu, 2012: Influence of the Kuroshio in the East China Sea on the early summer (baiu) rain. J. Climate, 25, 6627–6645, https://doi.org/10.1175/JCLI-D-11-00727.1.
Schneider, N., 2020: Scale and Rossby number dependence of observed wind responses to ocean-mesoscale sea surface temperatures. J. Atmos. Sci., 77, 3171–3192, https://doi.org/10.1175/JAS-D-20-0154.1.
Small, R. J., and Coauthors, 2008: Air–sea interaction over ocean fronts and eddies. Dyn. Atmos. Oceans, 45, 274–319, https://doi.org/10.1016/j.dynatmoce.2008.01.001.
Smith, B. L., S. E. Yuter, P. J. Neiman, and D. E. Kingsmill, 2010: Water vapor fluxes and orographic precipitation over Northern California associated with a landfalling atmospheric river. Mon. Wea. Rev., 138, 74–100, https://doi.org/10.1175/2009MWR2939.1.
Song, Q., P. Cornillon, and T. Hara, 2006: Surface wind response to oceanic fronts. J. Geophys. Res., 111, C12006, https://doi.org/10.1029/2006JC003680.
Sugimoto, S., K. Aono, and S. Fukui, 2017: Local atmospheric response to warm mesoscale ocean eddies in the Kuroshio–Oyashio confluence region. Sci. Rep., 7, 11871, https://doi.org/10.1038/s41598-017-12206-9.
Sugimoto, S., B. Qiu, and A. Kojima, 2020: Marked coastal warming off Tokai attributable to Kuroshio large meander. J. Oceanogr., 76, 141–154, https://doi.org/10.1007/s10872-019-00531-8.
Tachibana, Y., T. Iwamoto, M. Ogi, and Y. Watanabe, 2004: Abnormal meridional temperature gradient and its relation to the Okhotsk high. J. Meteor. Soc. Japan, 82, 1399–1415, https://doi.org/10.2151/jmsj.2004.1399.
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, 165–216.
Takahashi, H. G., S. A. Adachi, T. Sato, M. Hara, X. Ma, and F. Kimura, 2015: An oceanic impact of the Kuroshio on surface air temperature on the Pacific coast of Japan in summer: Regional H2O greenhouse gas effect. J. Climate, 28, 7128–7144, https://doi.org/10.1175/JCLI-D-14-00763.1.
Takatama, K., S. Minobe, M. Inatsu, and R. J. Small, 2012: Diagnostics for near-surface wind convergence/divergence response to the Gulf Stream in a regional atmospheric model. Atmos. Sci. Lett., 13, 16–21, https://doi.org/10.1002/asl.355.
Tanimoto, Y., T. Kanenari, H. Tokinaga, and S.-P. Xie, 2011: Sea level pressure minimum along the Kuroshio and its extension. J. Climate, 24, 4419–4434, https://doi.org/10.1175/2011JCLI4062.1.
Thom, E. C., 1957: A new concept for cooling degree days. Air Cond. Heat. Vent., 54, 73–80.
Thom, E. C., 1959: The discomfort index. Weatherwise, 12, 57–61, https://doi.org/10.1080/00431672.1959.9926960.
Tokinaga, H., Y. Tanimoto, M. Nonaka, B. Taguchi, T. Fukamachi, and S.-P. Xie, 2006: Atmospheric sounding over the winter Kuroshio Extension: Effect of surface stability on atmospheric boundary layer structure. Geophys. Res. Lett., 33, L04703, https://doi.org/10.1029/2005GL025102.
Tokinaga, H., Y. Tanimoto, S.-P. Xie, T. Sampe, H. Tomita, and H. Ichikawa, 2009: Ocean frontal effects on the vertical development of clouds over the western North Pacific: In situ and satellite observations. J. Climate, 22, 4241–4260, https://doi.org/10.1175/2009JCLI2763.1.
Vazquez-Cuervo, J., B. Dewitte, T. M. Chin, E. M. Armstrong, S. Purca, and E. Alburqueque, 2013: An analysis of SST gradients off the Peruvian coast: The impact of going to higher resolution. Remote Sens. Environ., 131, 76–84, https://doi.org/10.1016/j.rse.2012.12.010.
Wakabayashi, S., and R. Kawamura, 2004: Extraction of major teleconnection patterns possibly associated with the anomalous summer climate in Japan. J. Meteor. Soc. Japan, 82, 1577–1588, https://doi.org/10.2151/jmsj.82.1577.
Wallace, J. M., T. P. Mitchell, and C. Deser, 1989: The influence of sea surface temperature on sea surface wind in the eastern equatorial Pacific: Seasonal and interannual variability. J. Climate, 2, 1492–1499, https://doi.org/10.1175/1520-0442(1989)002<1492:TIOSST>2.0.CO;2.
Wu, R., and J. L. Kinter III, 2010: Atmosphere–ocean relationship in the midlatitude North Pacific: Seasonal dependence and east–west contrast. J. Geophys. Res., 115, D06101, https://doi.org/10.1029/2009JD012579.
Xie, S.-P., J. Hafner, Y. Tanimoto, W. T. Liu, H. Tokinaga, and H. Xu, 2002: Bathymetric effect on the winter sea surface temperature and climate of the Yellow and East China Seas. Geophys. Res. Lett., 29, 2228, https://doi.org/10.1029/2002GL015884.
Xu, H., H. Tokinaga, and S.-P. Xie, 2010: Atmospheric effects of the Kuroshio large meander during 2004–05. J. Climate, 23, 4704–4715, https://doi.org/10.1175/2010JCLI3267.1.
Xu, H., M. Xu, S.-P. Xie, and Y. Wang, 2011: Deep atmospheric response to the spring Kuroshio Current over the East China Sea. J. Climate, 24, 4959–4972, https://doi.org/10.1175/JCLI-D-10-05034.1.
Xu, M., H. Xu, and H. Ren, 2018: Influence of Kuroshio SST front in the East China Sea on the climatological evolution of Meiyu rainband. Climate Dyn., 50, 1243–1266, https://doi.org/10.1007/s00382-017-3681-2.
Yanai, M., and T. Tomita, 1998: Seasonal and interannual variability of atmospheric heat sources and moisture sinks as determined from NCEP–NCAR reanalysis. J. Climate, 11, 463–482, https://doi.org/10.1175/1520-0442(1998)011<0463:SAIVOA>2.0.CO;2.
Yanai, M., S. Esbensen, and J. H. Chu, 1973: Determination of bulk properties of tropical cloud clusters from large-scale heat and moisture budgets. J. Atmos. Sci., 30, 611–627, https://doi.org/10.1175/1520-0469(1973)030<0611:DOBPOT>2.0.CO;2.
Yasunaka, S., and K. Hanawa, 2006: Interannual summer temperature variations over Japan and their relation to large-scale atmospheric circulation field. J. Meteor. Soc. Japan, 84, 641–652, https://doi.org/10.2151/jmsj.84.641.