1. Introduction
The wintertime mean atmospheric circulation over the Far East and northwestern (NW) Pacific is characterized by a low-level northwesterly monsoonal flow between the Siberian high and Aleutian low and also by an extremely intensified westerly jet aloft. With this steady monsoonal flow, this region marks the strongest poleward heat transport over the Northern Hemisphere in the climatological mean (van Loon and Williams 1979; Higuchi et al. 1991; see also Fig. 1a). Over this region, the near-surface meridional temperature gradients are extremely high in midlatitudes (Fig. 1e), where the cold, dry monsoonal air encounters a warm air mass to the south. In addition, an abundant supply of heat and moisture occurs from the warm ocean surface to the monsoonal air (e.g., Esbensen and Kushnir 1981; see also Fig. 1g). These two effects sustain high lower-tropospheric baroclinicity (Hoskins and Valdes 1990; Nakamura 1992), which feeds migratory baroclinic eddies to form a well-defined storm track1 downstream (Blackmon et al. 1977, 1984; Wallace et al. 1988; see also Fig. 1b). It is marked with a belt of local maxima in precipitation almost across the Pacific basin (Xie and Arkin 1997; Trenberth and Guillemot 1998; see also Fig. 1h). These migratory eddies act to relax the temperature gradients by systematically transporting sensible heat to higher latitudes (e.g., Blackmon et al. 1977; Nakamura 1992; see also Fig. 1b).
It became apparent that the east Asian winter monsoon and associated meridional heat transport both undergo substantial modulations with interannual timescales (van Loon and Williams 1979; Higuchi et al. 1991; Zhang et al. 1997). It has now started to be realized that the winter monsoon is modulated also on decadal timescales (Koide and Kodera 1999; Nakamura and Yamagata 1999), with its weakening since the late 1980s as a distinct example (Watanabe and Nitta 1999). These modulations could potentially alter the activity of migratory disturbances along the storm track through the associated changes in the near-surface baroclinicity and jet intensity aloft over the Far East, the entrance region of the Pacific storm track. The anomalous growth of the eddies in the storm track entrance may influence the wintertime-mean circulation downstream through anomalous eddy vorticity flux convergence (Peng and Whitaker 1999). Anomalies in the monsoon and storm track may also accompany significant changes in the hydrological cycle over the North Pacific through anomalous evaporation from the warm ocean surface and precipitation along the storm track as well. Therefore, the interannual and decadal modulations in the winter monsoon and Pacific storm track could have significant climatic implications.
The relationship is, by nature, interactive between the activity of the eddies migrating along a storm track and the mean flow in which the eddies are embedded. In fact, stationary anomalies in the basic flow are subject to the continual feedback forcing from those eddies (e.g., Lau 1988; Lau and Nath 1991). Still, a significant part of the interaction may be understood at least qualitatively, once anomalies in the basic flow are specified. Lau (1988) showed that the observed anomalous behavior of the major Northern Hemisphere storm tracks, in general, can be interpreted reasonably well in view of how an anomalous basic flow reorganizes the storm track. A linear stability analysis applied by Robertson and Metz (1989) to several anomalous basic flows suggests that the reorganizing effect by an anomalous mean flow may be interpreted reasonably well within the framework of the linear theory of baroclinic instability. The effect includes the modification of the mean baroclinicity and the steering effect acting on the synoptic-scale eddies. Interpretation of the relationship observed between changes in the Pacific storm track and east Asian monsoon, however, may not necessarily be straightforward. This is because not only of the presence of the feedback forcing from the storm track, but also of the presence of a complex relationship between mean baroclinicity and eddy amplitude inherent to the Pacific storm track (Nakamura 1992). This relationship is manifested as a striking midwinter minimum in the climatological-mean storm track activity in spite of locally enhanced mean-flow baroclinicity (Nakamura 1992). This phenomenon has been reproduced in general circulation models (GCMs; Christoph et al. 1997; Zhang and Held 1999) and even in a linear stochastic model (Zhang and Held 1999). Very recently, Chang (2001) investigated seasonal and interannual variations in the Pacific storm track activity based on a GCM simulation and by comparing observational circulation statistics between a couple of contrasting winter seasons. He found that fairly systematic changes in the eddy structure and group velocity are important in the interannual variations. He also found that fall–winter differences in the precipitation processes associated with individual cyclones are more important in the particular seasonal cycle in the Pacific storm track.
In the present study, we document the interannual and (quasi) decadal changes in the storm track activity observed over the NW Pacific for 17 recent years. We also document the associated changes in the hydrological cycle over the North Pacific basin. As shown in the previous studies mentioned above, this period includes a profound wintertime climatic change over the Far East and North Pacific, associated with the weakening of the monsoonal flow and upper-tropospheric jet that occurred in the late 1980s. Specifically, we are interested in what kind of changes occurred in the Pacific storm track in its activity and feedback forcing over its entrance region and farther downstream as well in association with the changes in the monsoonal flow. We are also interested in any indication of the associated changes, as ensemble, in the structure of the eddies migrating along the storm track. In specific, we are interested in zonal phase alignment between eddy velocity components and temperature fluctuations, which could influence the efficiency for the eddies in converting energy from the mean flow for their growth. A brief summary of the present study is found in Nakamura and Izumi (1999).
2. Data and analysis procedures
In the present study, we use 6-hourly global circulation data based on the National Centers for Environmental Prediction–National Center for Atmospheric Research (NCEP–NCAR) reanalyses (Kalnay et al. 1996) from 1979 (autumn) to 1995 (summer). The data quality should be better for this period than for the earlier period, because of satellite-derived information available for the reanalysis scheme.
As in Nakamura and Wallace (1990) and Nakamura (1992), fluctuations associated with migratory synoptic-scale disturbances whose time periods are typically shorter than a week have been extracted by means of digital high-pass filtering with a half-power cutoff period of 8 days.2 In recognition of the baroclinic nature of those disturbances particularly in the lower troposphere, an instantaneous storm track activity at a particular location is measured as the 8-day low-pass-filtered meridional heat flux (
A stationary component that includes the mean westerly jet, low-level monsoonal circulation, and associated vertical motion and thermal fields has been extracted through 31-day averaging. In this study the meridional heat transport due to the persistent, stationary eddies including the monsoonal flow is evaluated at each location as the product of the 850-mb temperature and meridional velocity (
In our brief examination of the hydrological cycle over the North Pacific, we use not only the precipitation data based on the NCEP–NCAR reanalyses but also monthly data taken from the so-called Climate Prediction Center (CPC) Merged Analysis of Precipitation dataset (CMAP; Xie and Arkin 1997). In constructing the dataset, rain gauge measurements over land were combined with estimates over the oceans based on infrared and microwave radiation measurements by satellites, while the data from the NCEP–NCAR reanalyses were substituted for data missing grids mainly over the Arctic and southern oceans. Hence, the CMAP data are believed to bear higher credibility than the reanalysis precipitation data themselves that depend on model parameterization schemes (Trenberth and Guillemot 1998). Likewise, in our examination of turbulent fluxes of sensible and latent heat over the ocean surface, we use the reanalysis products and estimates based on in situ ship measurements in the Comprehensive Ocean–Atmosphere Data Set (COADS) archive as well. The former depends on the parameterization scheme of the surface boundary layer in the reanalysis, whereas the latter estimation suffers from insufficient time intervals in measurements that are likely to result in the underestimation of the contribution from submonthly transients to the fluxes.
The climatological mean of a given variable including the eddy heat fluxes is defined as the 16-yr mean of the 31-day moving-averaged value for each calendar day. Anomalies associated with the interannual variability are then represented as the 31-day averaged deviations from the climatological mean for a given day, and the overall strength of the variability is measured as its standard deviation for a given calendar day. Bimonthly mean anomalies for January and February of the satellite precipitation and COADS-based surface heat fluxes were simply calculated from their monthly anomalies, whereas a bimonthly anomaly field for any variable based on the reanalyses was constructed as the simple algebraic mean of the two 31-day running mean anomaly fields, one for the mid-January and the other for mid-February.
3. Interannual variability in the storm track and monsoon activities
Our main analysis domain is limited within the Far East and the NW Pacific (20°–60°N, 100°E–180°), where both the northwesterly monsoonal flow (
To elucidate the typical seasonal march and interannual variability in the monsoonal heat transport and storm track activity, the climatological means and standard deviations of the 850-mb
In order to identify the most prevailing interannual variability in the midwinter storm track activity within the analysis domain (Fig. 1d), a conventional empirical orthogonal function (EOF) analysis was applied to bimonthly (January–February) anomalies of
It is hinted in Figs. 3a and 4a that the relationship between the anomalous westerlies and anomalous storm track activity is different between the NW and NE Pacific. This difference is more apparent in the corresponding maps for the upper troposphere (Figs. 3b and 4b). Over the NE Pacific, the storm track activity tends to be deflected poleward at the upper and lower levels in the presence of stationary anticyclonic anomalies in midlatitudes. In the situation shown in Figs. 4a and 4b, the axis of the mean westerlies that steer synoptic-scale eddies is also shifted poleward, and the mean baroclinicity is anomalously high and low along the northern and southern flanks of the anticyclonic anomalies, respectively, whose vertical structure is approximately equivalent barotropic. In the upper troposphere, another significant anomaly in the storm track activity is found to the southwest of Hawaii (Fig. 3b), where the anomalous seasonal wind tends to be westerly. The relationship between the anomalous storm track activity and stationary anomalies over the NE Pacific is in agreement with Lau (1988), and it is understandable in the framework of linear theory of baroclinic instability as analyzed in Robertson and Metz (1989). Over the NW Pacific, in contrast, the storm track activity tends to increase in spite of the anomalous easterlies in the upper troposphere (Figs. 3 and 4b) and the reduced lower-tropospheric baroclinicity along the storm track (Fig. 4c). This result is in good agreement with the empirical relationship between baroclinic eddy amplitude and mean westerly speed obtained for the NW Pacific by Nakamura (1992, his Figs. 10a and 12a) and in a GCM simulation by Carillo et al. (2000). The relationship found between the anomalous storm track activity and basic-state anomalies over the NW Pacific cannot be interpreted in the framework of the linear theory of baroclinic instability.
4. Interannual variability in poleward heat transport over the Far East
A striking feature found in comparison between Figs. 3a and 4d is that the enhanced heat transport by migratory eddies (
In recognition of some ambiguities in defining a local poleward heat flux associated with the stationary eddies (or monsoonal flow), it is necessary to confirm the above findings from a different viewpoint. Specifically, for each midwinter season we evaluated the local heating rate due to the temperature advection associated with the 850-mb seasonal-mean monsoonal flow (−
5. Modulated seasonal march
In this section, seasonal dependency of the out-of-phase relationship between interannual fluctuations in meridional heat transport associated with the Pacific storm track and east Asian winter monsoon is highlighted in their composite latitude–season sections. Our selection of winters used for the compositing was based on the PC1 time series for midwinter
In the composite maps of the 850-mb
The seasonal march in the mean state is shown in the composites of U250 (Fig. 10) and ΔU700−1000 (Fig. 11). These composites indicate that both the upper-level jet and the associated lower-tropospheric baroclinicity over the NW Pacific tend to be stronger significantly in the negative winters than in the positive winters. At the same time, the jet and baroclinic zone tend to be meridionally narrower in the negative winters, which is apparent in the difference in their poleward extent (in Figs. 10a,b and 11a,b) and in the dipolar patterns in their anomalies (in Figs. 10c and 11c). These changes in the jet and baroclinicity occurred in conjunction with more enhanced southward spreading of a cold air mass in the negative winters than in the positive winters (Fig. 8). It is evident in Fig. 7 that the storm track axis tends to be shifted southward in the negative winters, relative to the situation in the positive winters. This shift is much more apparent than the corresponding shifts in the axes of the jet and baroclinic zone. Under the enhanced monsoonal flow, the storm track tends to be closer to the jet axis and low-level baroclinic zone whereas the activity of the eddies is strongly suppressed.
In the PC1 time series of the storm track activity shown in Fig. 5a, the decadal variability, which changed its polarity in the late 1980s, is apparent even without any smoothing imposed. Correspondingly, all of the five “positive winters” listed above are sampled from the period of the late 1980s to the mid-1990s, whereas 4 out of the 5 negative winters are in the early and mid-1980s. Though slightly less pronounced, the same kind of distinctions as above in the Pacific storm track activity and in the mean state were found, when the winter of 1994/95 was replaced with that of 1984/85 in the cluster of the negative winters for the compositing.3 It is suggested that substantial modulations occurred in the seasonal march of the Pacific storm track activity, in conjunction with the recent decadal-scale weakening tendency in the east Asian winter monsoon. In fact, the midwinter intensity of the Siberian high varied predominantly on decadal scales during the recent period, with a rapid weakening around 1987 (Fig. 5b). Though masked by the dominant year-to-year fluctuations associated mainly with ENSO events, a decadal-scale tendency is noticeable also in the intensity of the Aleutian low (Fig. 5b). It tended to be stronger until the late 1980s and weaker since then. The decadal modulation in the storm track activity is also evident in Fig. 12, where the 850-mb
6. Possible physical mechanisms for the observed storm track changes
As shown in the preceding section, the observed midwinter suppression (enhancement) in the activity of the Pacific storm track occurred in spite of the intensified (relaxed) low-level baroclinicity and upper-level jet. The intensified baroclinicity favors the amplification of baroclinic eddies, but the intensified jet may counteract it by advecting the eddies too fast away from a zonally confined baroclinic zone (e.g., Lin and Pierrehumbert 1993). Nakamura (1992) speculated that the latter effect may contribute to the midwinter minimum in the climatological seasonal march in the Pacific storm track activity. In order to assess whether the interannual variations in that activity shown in the preceding section can be attributed to the excessive advective effect, we evaluated a parameter dG defined as dG = [f(ϕs)
The two components of dG were evaluated separately for each of the two sets of winters on the basis of the 5-yr composites of
Putting this possible overestimation aside, we find in Fig. 14 that G differs only about 5% between the composites for the strongest and weakest
Nakamura (1992) showed that in climatology the correlation between the 850-mb temperature and meridional wind velocity in the high-pass-filtered fields tends to be lower in midwinter than in fall and spring. Similar changes in the temperature–velocity correlation associated with baroclinic eddies were found by Chang (2001) in his analysis of a GCM simulation and the observed data. As discussed in detail in the appendix, the velocity–temperature correlation in the high-pass-filtered data is an indicator of their phase relationship as an ensemble of migratory eddies. The correlation statistics are unlikely to influenced by the number of synoptic-scale disturbances or their amplitude observed within a particular season. One may wonder whether the observed changes in the midwinter storm activity as in Fig. 7 are attributable to the significant changes in eddy structure. In order to examine if that was really the case, a correlation analysis was applied at each grid point between the 8-day high-pass-filtered time series of the 850-mb temperature and meridional wind velocity for the January–February period separately for the five positive and negative winters defined in section 5. The same local correlation analysis was applied between the high-pass-filtered time series of the 500-mb temperature and vertical P velocity. The former and latter correlations are hereafter referred to as the 850-mb T–υ correlation and 500-mb T–ω correlation, respectively. For baroclinically growing eddies, the former correlation should be positive but the latter should be negative because of the sign convention.
Figure 15 contrasts each of the two types of correlation between the positive and negative winters. The 850-mb T–υ correlation is higher in the positive winters than in the negative winters. In the former winters the correlation exceeds 0.7 in many locations along the storm track (Fig. 15a), while in the latter winters it is generally below 0.7 and even below 0.6 in some locations along the storm track (Fig. 15b). The 500-mb T–ω correlation also exhibits similar differences. The negative correlation is −0.4 or even stronger throughout the Pacific storm track in the positive winters (Fig. 15c), whereas in the negative winters the correlation is substantially weaker in general and it is even weaker than −0.2 around Japan (Fig. 15d). From an energetic viewpoint, the stronger T–υ correlation means more efficient conversion of available potential energy from the mean westerlies to baroclinic disturbances that tend to be more baroclinic than in the climatological situation (see the appendix). Likewise, the stronger T–ω correlation means more efficient conversion from available potential energy to kinetic energy within those disturbances. The results shown in Fig. 15 suggest that the synoptic-scale eddies tended to lose their optimum structure for their baroclinic growth in the presence of the excessively strong westerlies in the negative winters. These results are consistent with those in the previous studies mentioned above.
7. Influence upon the hydrological cycle and atmosphere–ocean heat exchange
In order to examine whether the recent decadal variations in the storm track and monsoon activities over the NW Pacific as documented in the preceding sections were accompanied by any significant changes in the hydrological budget or atmosphere–ocean heat exchange over the North Pacific, January–February mean fields of precipitation, water vapor, and turbulent heat fluxes were linearly regressed upon PC1 of 850-mb
The decadal changes in the midwinter atmospheric circulation and storm track activity are accompanied by a coherent pattern of anomalous latent heat release from the ocean surface (Fig. 16a). Weakened spills of cold, dry air from the Asian continent in the winter of the weak monsoon suppress latent heat release by 5%–10% over the warm Kuroshio, its extension, and branches. Suppressed latent heat release farther to the east is due to the reduced surface westerlies associated with the weakened Aleutian low. The weakened low also reduces the transport of relatively warm and moist air from the south along the west coast of Canada, which acts to increase the evaporation from the ocean surface by 10%–15%. In the subtropics, situated to the south of the anticyclonic anomalies covering over the midlatitude North Pacific, the latent heat release from the ocean surface is enhanced by as much as 15% under the intensified northeasterly trades. Sensible heat flux anomalies are distributed in a manner very similar to that of the latent heat flux anomalies (Fig. 16b), but the former is about a half in magnitude of the latter in midlatitudes and even about 30% in the subtropics. Again, the above argument signifies a typical situation since the late 1980s, and the sign of each anomaly should be reversed in discussing the counterpart for the early to mid-1980s. These results based on the surface flux fields of the NCEP–NCAR reanalysis data are reproduced reasonably well on the basis of anomalous flux fields estimated from the COADS archive.
With respect to the hydrological cycle, the increased rainfall over the subtropics that occurs in the proxy of the region of enhanced evaporation implies the availability of a local moisture supply. Indeed, no significant anomalous convergence of an atmospheric moisture flux is found in that subtropical region (Fig. 16d). In midlatitudes, the increased precipitation along the intensified storm track in the region of the suppressed evaporation over the Kuroshio and the opposite situation over the NE Pacific yield the anomalous net moisture sink and source in these respective regions. Accordingly, we observe significant anomalous convergence and divergence of the moisture flux along the western half of the Pacific storm track and over the extreme NE Pacific, respectively. Correspondingly, a clear signal of the anomalous westward transport of moisture is found along the storm track across the Pacific. Though less pronounced, another anomalous moisture transport into the entrance of the intensified storm track is present from the subtropics. Figure 16d represents a typical situation since the late 1980s, and those anomalous moisture transports tended to be reversed in the first half of the 1980s. The amount of the anomalous zonal moisture transport is such that it could compensate most of the net anomalous moisture deficit around Japan as inferred from Figs. 16a and 16c if the transport as in Fig. 16d occurs within the lowest 3 km of the atmosphere. It is concluded that the observed modulation in the storm track activity over the North Pacific in the presence of the decadal weakening of the winter monsoon and Aleutian low was accompanied by significant changes in precipitation, evaporation, moisture transport, and atmosphere–ocean heat exchange over the entire North Pacific. It should be kept in mind that the anomalous moisture transport shown in Fig. 16d is probably more or less overestimated, because the precipitation decrease along the climatological-mean storm track in the NE Pacific since the late 1980s appears to be somewhat overestimated in the reanalyses compared to the CMAP data.
8. Summary and discussion
We have shown that a profound modulation in the seasonal march in the NW Pacific storm track activity occurred concomitantly with the decadal weakening of the east Asian winter monsoon (Siberian high) and Aleutian low during the 1980s and 1990s. The most notable signature of this modulation is the tendency that the midwinter minimum in storm track activity became less apparent in the second half of this period than in the first half, in association with a increasing tendency in the midwinter activity. This enhancement of the storm track activity occurred despite the reduced mean-flow baroclinicity and jet intensity aloft, which is opposed to linear theory of baroclinic instability and is analogous to the eddy–mean-flow relationship in the midwinter suppression within the climatological-mean seasonal cycle (Nakamura 1992). At this stage, no solid argument can be made on what mechanism caused that particular decadal enhancement in the midwinter Pacific storm track activity. Yet, our analysis suggests that, as the excessively strong westerlies weakened, the eddies changed their structure so as to be more efficient in converting the mean-flow available potential energy into their kinetic energy for their growth. Our analysis also implies that the weakening of the Pacific jet in the late 1980s prolonged the residence time for growing eddies within the zonally confined baroclinic zone beneath the jet. This particular effect favored the amplification of those eddies and seemingly dominated over the effect of the reduced mean baroclinicity, although this effect appears to be of secondary importance. Our findings are in good agreement with Chang (2001), who showed that the reduced energy conversion from the mean-flow available potential energy and increased eastward group velocity both contributed to the suppression of downstream eddy growth along the Pacific storm track in a winter when the Pacific jet is particularly strong.
This observed relationship between the anomalous storm track activity and mean baroclinicity shown in this paper is found unique to the NW Pacific, as the corresponding relationship is found different for another major storm track over the NW Atlantic. The same analysis as in section 3 applied to the January–February anomaly fields within (30°–60°N, 100°–30°W) reveals that the strongest interannual variability in
As mentioned earlier, the presence of the feedback forcing from eddies migrating along the storm track renders it difficult to identify any causal relationship between changes in the mean flow and eddy activity as obtained through our analysis. Over the central and eastern Pacific, the difference map of ω500 indicates a clear signature of the enhancement of a thermally indirect cell to the north of the mean storm track axis (Fig. 18a), which is consistent with the northward deflection of the storm track extracted in the first EOF (Fig. 3a). In addition, the westerly acceleration through eddy vorticity flux tended to be significantly stronger just to the north of the storm track axis and weaker over the subtropics in the positive winters than in the negative winters (Fig. 18b). This upper-level barotropic forcing includes a component that is in phase with the observed U250 anomalies, and it acts to vary the low-level wind accordingly via ageostrophic circulation. Thus the system appears to be interactive over the central and eastern North Pacific, and hence the anomalous storm activity was likely to contribute to the observed anomalous Aleutian low and associated anomalies aloft as in Figs. 4a and 4b. Over the midlatitude western Pacific and Far East, in contrast, no significant difference was found in the barotropic feedback forcing from the anomalous storm track activity (Fig. 18b). Furthermore, a significant anomaly pattern evident over the midlatitude Far East in the difference map of ω500 is in thermally direct sense (Fig. 18a), which cannot be attributed to the enhanced storm track activity in the positive winters. Therefore, the anomalies in temperature and vertical motion observed over the Far East are unlikely to be generated in direct response to the feedback forcing from the local storm track. One may argue that quasi-stationary anomalies generated in response to the anomalous feedback forcing from migratory eddies in the core and exit regions of the Pacific storm track may retrograde slowly and possibly give rise to the temperature anomalies in the entrance region of the storm track over the Chinese continent as in Fig. 4c. However, no such indication of retrogression was found in a Hovmöller diagram of 8-day low-pass-filtered 250-mb height anomalies for the January–February period for any of the five “positive” or five “negative” winters (not shown). As discussed later, the anomalous vertical motion over the Far East might occur in association with an upper-level stationary wave train propagating across Eurasia.
Nakamura and Yamagata (1999) pointed out that 2 of the 3 dominant patterns of the decadal variability in the NW Pacific wintertime SST changed their polarities in the later half of the 1980s. One of them is what may be called a “midlatitude pattern,” characterized by SST anomalies along the subarctic front associated with the anomalous Aleutian low and the PNA teleconnection pattern aloft. Their polarities changed around 1988/89. The other represents SST anomalies in the East China Sea, which is considered to be generated in thermal response to the decadal modulation of the east Asian winter monsoon. The associated atmospheric pattern changed its polarity around 1986/87,4 just when the weakening of the Siberian high occurred (Fig. 5b). Our analysis seems to capture some systematic changes in the Pacific storm track activity associated with the polarity changes of those two patterns of the decadal climate variability. In fact, the unsmoothed time series of PC1 of
Figure 16 indicates that the decadal-scale modulation observed in the east Asian winter monsoon and the associated modulation in the storm track activity influenced the atmosphere–ocean heat exchange and the hydrological cycle over the entire North Pacific as well. The weakening of the winter monsoon in the late 1980s caused the reduced evaporation over the extreme NW Pacific around Japan, which contributed to the observed warming in this maritime region (Nakamura and Yamagata 1999). The weakening of the Aleutian low contributed to the reduction and enhancement in surface evaporation over the midlatitude North Pacific and off the west coast of North America, respectively, which reinforced the warming and cooling of SST observed in the respective regions (Nakamura et al. 1997; Nakamura and Yamagata 1999). The enhancement and suppression in the storm track activity observed over the NW and NE midlatitude Pacific, respectively, were accompanied by the increased and decreased precipitation in these respective regions (Fig. 16c). These decadal anomalies in precipitation may be influential on the ocean circulation through changing density structure in the upper ocean. In fact, Lukas (2001) observed a marked increase in salinity in 1990/91 at the main thermocline level and above (between 150 and 350 m in depth) near the Oahu Island of Hawaii. Isopycnal surfaces crossing those depths at that particular location are ventilated within the oceanic subtropical frontal zone located to the north. This observed salinity change seems consistent with the below-normal precipitation during the first half of the 1990s over the oceanic subtropical frontal zone (Lukas 2001; see also Fig. 16c), allowing for a time lag with which the salinity increase thus generated in the mixed layer over the frontal zone was subducted and then advected southward by a geostrophic current. It should be mentioned that the characteristics of deepest mixing also depend on intensity and frequency of strongest storms in each winter season. Hence, the storm track variations as shown in the present study may be important in the oceanic variability not only through changing the net supply of freshwater at the ocean surface but also through altering the strength of subduction.
Although the aforementioned profound signatures were found in the present study, our analysis is based on a data period that is obviously too short to capture the robust characteristics of the decadal-scale storm track variability. Therefore, our analysis should be regarded as a case study concerning a single transition event associated with the decadal weakening of the winter monsoon that occurred in the late 1980s. An attempt is currently under way to examine whether any coherent pattern as identified in the 1980s and 1990s can be extracted for the wintertime North Pacific from the entire 40-yr period of the NCEP–NCAR reanalyses. A preliminary result suggests that the leading EOF of the variability in the NW Pacific storm track activity in midwinter and associated seasonal-mean anomalies in the wind and thermal fields are very similar to those presented in the current study.
Acknowledgments
We thank Dr. Edmund (K.-M.) Chang for his comments that were very helpful in interpreting the results of our analysis. We also thank Drs. Petros Ioannou, Roger Lukas, and Noboru Nakamura for their encouragement and stimulating discussion; Mr. Akihiko Sinpo and Dr. Koutarou Takaya for processing the NCEP–NCAR reanalysis data. Sound criticism and insightful suggestions given by the two anonymous referees on the earlier version of the paper led to its substantial improvement.
REFERENCES
Blackmon, M. L., J. M. Wallace, N-C. Lau, and S. L. Mullen, 1977: An observational study of the Northern Hemisphere wintertime circulation. J. Atmos. Sci., 34 , 1040–1053.
Blackmon, M. L., Y-H. Lee, and J. M. Wallace, 1984: Horizontal structure of 500-mb height fluctuations with long, intermediate and short time scales. J. Atmos. Sci., 41 , 961–979.
Carillo, A., P. M. Ruti, and A. Navarra, 2000: Storm tracks and zonal mean flow variability: A comparison between observed and simulated data. Climate Dyn., 16 , 219–228.
Chang, E. K. M., 1993: Downstream development of baroclinic waves as inferred from regression analysis. J. Atmos. Sci., 50 , 2038–2053.
Chang, E. K. M., . 2001: GCM and observational diagnoses of the seasonal and interannual variations of the Pacific storm track during the cold season. J. Atmos. Sci., 58 , 1784–1800.
Chang, E. K. M., and D. B. Yu, 1999: Characteristics of wave packets in the upper troposphere. Part I: Northern Hemisphere winter. J. Atmos. Sci., 56 , 1708–1728.
Charney, J. G., 1947: The dynamics of long waves in a baroclinic westerly current. J. Meteor., 4 , 135–163.
Christoph, M., U. Ulbrich, and P. Speth, 1997: Midwinter suppression of Northern Hemisphere storm track activity in the real atmosphere and in GCM experiments. J. Atmos. Sci., 54 , 1589–1598.
Eady, E. T., 1949: Long waves and cyclone waves. Tellus, 1 , 33–52.
Esbensen, S. K., and Y. Kushnir, 1981: The heat budget of the Global Ocean: An atlas based on estimates from surface marine observations. Climatic Research Institute Rep. 29, 27 pp. [Available from Department of Atmospheric Sciences, Oregon State University, Corvallis, OR 97331.].
Higuchi, K., C. A. Lin, A. Shabbar, and J. L. Knox, 1991: Interannual variability of the January tropospheric meridional eddy sensible heat transport in the Northern Hemisphere. J. Meteor. Soc. Japan, 69 , 459–472.
Hoskins, B. J., and P. J. Valdes, 1990: On the existence of storm tracks. J. Atmos. Sci., 47 , 1854–1864.
Kalnay, E., and Coauthors. 1996: The NCEP/NCAR 40-Year Reanalysis Project. Bull. Amer. Meteor. Soc., 77 , 437–471.
Koide, H., and K. Kodera, 1999: A SVD analysis between the winter NH 500-hPa height and surface temperature fields. J. Meteor. Soc. Japan, 77 , 47–61.
Lau, N-C., 1988: Variability of the observed midlatitude storm tracks in relation to low-frequency changes in the circulation pattern. J. Atmos. Sci., 45 , 2718–2743.
Lau, N-C., and M. J. Nath, 1991: Variability of the baroclinic and barotropic transient eddy forcing associated with monthly changes in the midlatitude storm tracks. J. Atmos. Sci., 48 , 2589–2613.
Lin, S. J., and R. T. Pierrehumbert, 1993: Is the midlatitude zonal flow absolutely unstable? J. Atmos. Sci., 50 , 505–517.
Lukas, R. B., 2001: Freshening of the upper pycnocline in the North Pacific subtropical gyre associated with decadal changes of rainfall. Geophys. Res. Lett., 28 , 3485–3488.
Nakamura, H., 1992: Midwinter suppression of baroclinic wave activity in the Pacific. J. Atmos. Sci., 49 , 1629–1641.
Nakamura, H., . 1993: Horizontal divergence associated with zonally isolated jet streams. J. Atmos. Sci., 50 , 2310–2313.
Nakamura, H., and J. M. Wallace, 1990: Observed changes in baroclinic wave activity during the life cycles of low-frequency circulation anomalies. J. Atmos. Sci., 47 , 1100–1116.
Nakamura, H., and T. Izumi, 1999: Out-of-phase relationship between the interannual fluctuations in the poleward heat transport by the East Asian winter monsoon and Pacific storm track. Preprints, 12th Conf. on Atmospheric and Oceanic Fluid Dynamics, New York, NY, Amer. Meteor. Soc., 139–142.
Nakamura, H., and T. Yamagata, 1999: Recent decadal SST variability in the Northwestern Pacific and associated atmospheric anomalies. Beyond El Niño: Decadal and Interdecadal Climate Variability, A. Navarra, Ed., Springer, 49–72.
Nakamura, H., G. Lin, and T. Yamagata, 1997: Decadal climate variability in the North Pacific during the recent decades. Bull. Amer. Meteor. Soc., 78 , 2215–2225.
Peng, S., and J. S. Whitaker, 1999: Mechanisms determining the atmospheric response to midlatitude SST anomalies. J. Climate, 12 , 1393–1408.
Robertson, A. W., and W. Mez, 1989: Three-dimensional linear instability of persistent anomalous large-scale flows. J. Atmos. Sci., 46 , 2783–2801.
Takaya, K., and H. Nakamura, 2001: A formulation of a phase-independent wave-activity flux for stationary and migratory quasi-geostrophic eddies on a zonally-varying basic flow. J. Atmos. Sci., 58 , 608–627.
Trenberth, K. E., and C. J. Guillemot, 1998: Evaluation of the atmospheric moisture and hydrological cycle in the NCEP/NCAR reanalyses. Climate Dyn., 14 , 213–231.
van Loon, H., and J. Williams, 1979: The association between latitudinal temperature gradient and eddy heat transport. Part II: Relationships between sensible heat transport by stationary waves and wind, pressure and temperature in winter. Mon. Wea. Rev., 108 , 604–614.
Wallace, J. M., and D. S. Gutzler, 1981: Teleconnections in geopotential height field during the Northern Hemisphere winter. Mon. Wea. Rev., 109 , 784–812.
Wallace, J. M., G-H. Lim, and M. L. Blackmon, 1988: Relationship between cyclone tracks, anticyclone tracks and baroclinic waveguides. J. Atmos. Sci., 45 , 439–462.
Watanabe, M., and T. Nitta, 1999: Decadal changes in the atmospheric circulation and associated surface climate variations in the Northern Hemisphere winter. J. Climate, 12 , 494–510.
Whittaker, L. M., and L. H. Horn, 1984: Northern Hemisphere extratropical cyclone activity for four mid-season months. J. Climatol., 4 , 297–310.
Xie, P., and P. A. Arkin, 1997: Global precipitation: 17-year monthly analysis based on gauge observations, satellite estimates and numerical model outputs. Bull. Amer. Meteor. Soc., 78 , 2539–2558.
Xie, S-P., H. Noguchi, and S. Matsumura, 1999: A hemispheric-scale quasi-decadal oscillation and its signature in northern Japan. J. Meteor. Soc. Japan, 77 , 573–582.
Zhang, Yi, K. R. Sperber, and J. S. Boyle, 1997: Climatology and interannual variation of the East Asian winter monsoon: Results from 1979–95 NCEP/NCAR reanalysis. Mon. Wea. Rev., 125 , 2605–2619.
Zhang, Yu, and I. M. Held, 1999: A linear stochastic model of a GCM's midlatitude storm tracks. J. Atmos. Sci., 56 , 3416–3435.
APPENDIX
Eddy Correlation Statistics
For simplifying our argument, suppose a pair of synoptic-scale cyclonic and anticyclonic eddies were lined up side by side in the zonal direction. We may regard these anomalies as wavy baroclinic disturbances passing a particular location over a time period of T (days) that corresponds to the wave period. The associated fluctuations at a given pressure level in the meridional velocity υ and potential temperature θ measured over the period T would yield the positive correlation coefficient rυθ between them. It is independent of their amplitudes and simply reflects the zonal phase relationship between the υ and θ fluctuations. In an ideal example, rυθ = coskx0 for the disturbances represented as υ(x, t) = υn cosk(x − ct) and θ(x, t) = θn cosk(x + x0 − ct), where kx0 denotes their phase shift in the direction of eddy propagation, a certain aspect of the eddy structure. The standard deviations in υ and θ are given by σn(υ) = υn/
From an energetic viewpoint, the rate of the local conversion of available potential energy from the seasonal mean state to the high-frequency eddies may be defined as GA = CAT/AT, where AT is available potential energy (APE) associated with transient eddies and CAT its conversion from the seasonal-mean flow to the eddies both of which were evaluated locally throughout a winter season. Here, we assume the zonal uniformity of the mean flow for simplicity. For the “positive” winters, GAP = −(∂Θ/∂y)P[υθ]P/[
Likewise, the rate of local conversion from eddy APE to the eddy kinetic energy (KT) may be defined as GK = CKT/KT, where CKT denotes that conversion. Then, for the positive winters, GKP ≈ −2h(p)RPσP(ω)σP(θ)/
Climatological-mean distribution of 850-mb meridional heat fluxes (K m s−1) associated with (a) standing eddies (
Citation: Journal of Climate 15, 14; 10.1175/1520-0442(2002)015<1855:IADMRO>2.0.CO;2
(a), (b) Seasonal march of the climatological-mean 850-mb
Citation: Journal of Climate 15, 14; 10.1175/1520-0442(2002)015<1855:IADMRO>2.0.CO;2
(a) The leading mode of the interannual variability in the 850-mb Jan–Feb
Citation: Journal of Climate 15, 14; 10.1175/1520-0442(2002)015<1855:IADMRO>2.0.CO;2
Jan–Feb linear regression maps of (a) 850-mb height (m; every 5, heavy lines for 20 and 40), (b) 250-mb height (m; every 10, heavy lines for ±50), (c) 850-mb temperature (K; every 0.2, heavy lines for ±1), and (d) 850-mb
Citation: Journal of Climate 15, 14; 10.1175/1520-0442(2002)015<1855:IADMRO>2.0.CO;2
(a) Normalized time series of the leading principal component (PC1) of the interannual variability in the 850-mb
Citation: Journal of Climate 15, 14; 10.1175/1520-0442(2002)015<1855:IADMRO>2.0.CO;2
(a) Latitudinal distribution of anomalous poleward heat fluxes (K m s−1) at the 850-mb level for Jan and Feb in association with the storm track activity (
Citation: Journal of Climate 15, 14; 10.1175/1520-0442(2002)015<1855:IADMRO>2.0.CO;2
Composite seasonal marches in the 850-mb
Citation: Journal of Climate 15, 14; 10.1175/1520-0442(2002)015<1855:IADMRO>2.0.CO;2
As in Fig. 7, but for the 850-mb monsoonal heat transport
Citation: Journal of Climate 15, 14; 10.1175/1520-0442(2002)015<1855:IADMRO>2.0.CO;2
Composite seasonal marches in the 250-mb storm track activity as measured by 250-mb Zenv which was sampled along the instantaneous storm track axis for the (a) five positive (1986/87, 1988/89, 1989/90, 1991/92, 1993/94) and (b) five negative (1980/81, 1982/83, 1983/84, 1985/86, 1994/95) winters, plotted in their time–longitude sections. Contour interval is 10 m, with heavy lines for 120 and shading between 100 and 110 m. The envelope function has been multiplied by a factor [sin(45°N)/sin(lat)], to mimic the amplitude in the geostrophic streamfunction. The sampling was performed within a 10° latitudinal band centered at the maximum 850-mb
Citation: Journal of Climate 15, 14; 10.1175/1520-0442(2002)015<1855:IADMRO>2.0.CO;2
As in Fig. 7, but for the 250-mb zonal wind velocity (U250). Contour intervals are 10 m s−1, with heavy lines for 50 in (a) and (b), and 2 m s−1, with heavy lines for −10 in (c)
Citation: Journal of Climate 15, 14; 10.1175/1520-0442(2002)015<1855:IADMRO>2.0.CO;2
As in Fig. 7, but for the ΔU700−1000. Contour intervals are 1 m s−1, with heavy lines for 5 in (a) and (b), and 0.2 m s−1, with heavy lines for ±1 (lines for 0 and ±0.2 are omitted) in (c). Negative contours are dashed
Citation: Journal of Climate 15, 14; 10.1175/1520-0442(2002)015<1855:IADMRO>2.0.CO;2
Year–month section showing interannual modulations in the seasonal march of the storm track activity over the NW Pacific. The activity is measured as the 850-mb
Citation: Journal of Climate 15, 14; 10.1175/1520-0442(2002)015<1855:IADMRO>2.0.CO;2
Composite seasonal marches in a parameter dG sampled along the instantaneous storm track axis for the (a) five positive (1986/87, 1988/89, 1989/90, 1991/92, and 1993/94) and (b) five negative (1980/81, 1982/83, 1983/84, 1985/86, 1994/95) winters, plotted in their time-longitude sections. The parameter is roughly proportional to the amplification rate baroclinic eddies undergo while traveling between a given grid interval (2.5° in longitude). Contour interval is 0.1. Values between 0.5 and 0.6 are indicated by hatching, between 0.6 and 0.7 by heavy shading, and between 0.4 and 0.5 by light shading. See text for details
Citation: Journal of Climate 15, 14; 10.1175/1520-0442(2002)015<1855:IADMRO>2.0.CO;2
Seasonal marches in G, i.e., the zonally integrated dG from 100°E to 180° along the storm track axis, for the positive (open circles) and negative (closed circles) winters that correspond to Figs. 13a and 13b, respectively. Tick marks along the abscissa indicate the first days of individual months
Citation: Journal of Climate 15, 14; 10.1175/1520-0442(2002)015<1855:IADMRO>2.0.CO;2
Correlation maps between the 8-day high-pass-filtered time series of temperature and meridional wind velocity at the 850-mb level in Jan and Feb, obtained separately for the five (a) positive and (b) negative winters, as indicated. Contour interval is every 0.1 from 0.4, with heavy lines for 0.6. Shaded lightly and heavily is where the correlation coefficient is between 0.6 and 0.7 and greater than 0.7, respectively. (c), (d) As in (a) and (b), but for correlation between temperature and vertical P velocity (ω) at the 500-mb level. Contour interval is every 0.1 from −0.1 to −0.6. Shaded lightly and heavily is where the correlation coefficient is between −0.4 and −0.5 and between −0.2 and −0.3, respectively. Values between −0.5 and −0.6 are indicated by hatching
Citation: Journal of Climate 15, 14; 10.1175/1520-0442(2002)015<1855:IADMRO>2.0.CO;2
Jan–Feb linear regression maps of (a) latent heat flux and (b) sensible heat flux from the surface. Contour interval is 5 W m−2, and maps are based on the NCEP–NCAR reanalyses. (c) As in (a) but for precipitation based on the CMAP data. Contour interval is 0.2 mm day−1. (d) As in (a) but for 850-mb irrotational moisture flux (arrows) and its divergence (dashed contours) and convergence (solid contours) based on the reanalyses. The scaling of arrows is given at the lower right corner. Contour interval is 0.1 g kg−1 day−1. In (a)–(d), negative contours are dashed and zero lines omitted. Shaded lightly and heavily is where the local correlation with PC1 exceeds the 90% and 95% confidence levels, respectively. These maps represent typical seasonal-mean anomalies when PC1 increases by a unit std dev
Citation: Journal of Climate 15, 14; 10.1175/1520-0442(2002)015<1855:IADMRO>2.0.CO;2
(a) The leading mode of the interannual variability in the 850-mb Jan–Feb
Citation: Journal of Climate 15, 14; 10.1175/1520-0442(2002)015<1855:IADMRO>2.0.CO;2
(a) Difference composite map for the Jan–Feb period of ω500, obtained by subtracting the composite for the five negative winters from the counterpart for the five positive winters. Contour interval is 0.01 Pa s−1, with dashed lines for negative values (anomalous uprising) and zero lines omitted. (b) As in (a), but for the 250-mb feedback forcing from the storm track measured as the westerly acceleration (∂U250/∂t) migratory transient eddies act to induce through their vorticity flux. Contour interval is 1 m s−1 day−1, with dashed lines for negative values (anomalous easterly acceleration) and zero lines omitted. In (a) and (b), the climatological-mean axis of the Pacific storm track at the 250-mb level is defined by the largest value of the mean Zenv at each meridian
Citation: Journal of Climate 15, 14; 10.1175/1520-0442(2002)015<1855:IADMRO>2.0.CO;2
As in Fig. 4, but for 500-hPa geopotential height. Contour interval is 20 m, with dashed lines for negative values, and zero lines omitted. Superimposed with arrows is the horizontal component of a wave activity flux defined by Takaya and Nakamura (2001) for stationary Rossby waves, which is parallel to the local group velocity. Scaling (m2 s−2) is indicated at the lower right corner. Sign of the height anomalies is such that they represent a typical situation in the late 1980s and early 1990s. The sign must be reversed in representing a typical situation in the early and mid-1980s, but the sign reversal is not necessary for the wave activity flux
Citation: Journal of Climate 15, 14; 10.1175/1520-0442(2002)015<1855:IADMRO>2.0.CO;2
In this study the term “storm track” signifies a region of strong high-frequency fluctuations associated with synoptic-scale baroclinic waves rather than the path of an individual cyclone. A storm track corresponds to a “baroclinic waveguide” in the terminology of Wallace et al. (1988). Although the exact location of a storm track thus defined is somewhat different from a cyclone path especially over the middle of an ocean basin, they exhibit a reasonable correspondence (cf. Whittaker and Horn 1984). Also, the entrance region of a storm track where the variance of high-pass-filtered quantity (or the covariance of such quantities) increases toward downstream corresponds to a region of frequent cyclogenesis defined from a synoptic analysis (cf. Whittaker and Horn 1984).
As in the previous studies cited above, we used a 4-pole tangent-Butterworth filter that yields no phase shift in the filtered time series. In the present study, the cutoff period was set to 8 days, rather than 6 days in those studies, in order to retain part of the variance near the lower end of the frequency domain associated with the baroclinic eddies. Change (1993) showed that the particular portion of the variance is important in accounting for wave packet–like behavior of those eddies.
The former is the winter of the fifth weakest storm track activity, and the latter barely missed being included in the five winters of the weakest storm track activity. See the time series of PC1 in Fig. 5a.
Since these two patterns of the decadal SST variability over the NW Pacific have been identified through a conventional EOF analysis by Nakamura and Yamagata (1999), the polarity change of the midlatitude pattern in 1988/89 was likely to be a coincidental sequel of that of the “monsoon pattern” in 1986/87.