Advanced Search
Article Contents

Changes in Mixed Layer Depth and Spring Bloom in the Kuroshio Extension under Global Warming


doi: 10.1007/s00376-015-5113-8

  • The mixed layer is deep in January-April in the Kuroshio Extension region. This paper investigates the response in this region of mixed layer depth (MLD) and the spring bloom initiation to global warming using the output of 15 models from CMIP5. The models indicate that in the late 21st century the mixed layer will shoal, and the MLD reduction will be most pronounced in spring at about 33°N on the southern edge of the present deep-MLD region. The advection of temperature change in the upper 100 m by the mean eastward flow explains the spatial pattern of MLD shoaling in the models. Associated with the shoaling mixed layer, the onset of spring bloom inception is projected to advance due to the strengthened stratification in the warming climate.
  • 加载中
  • Behrenfeld M. J., 2010: Abandoning Sverdrup's critical depth hypothesis on phytoplankton blooms. Ecology, 91( 4), 977- 989.10.1890/09-1207.120462113c5f371e7-a7e5-4650-8b94-16eb52e4fc427c62894363a33166d36eb6c64f43dbcahttp%3A%2F%2Fmed.wanfangdata.com.cn%2FPaper%2FDetail%2FPeriodicalPaper_PM20462113refpaperuri:(41c057236e3f33f8516d8c08279980fd)http://med.wanfangdata.com.cn/Paper/Detail/PeriodicalPaper_PM20462113The Critical Depth Hypothesis formalized by Sverdrup in 1953 posits that vernal phytoplankton blooms occur when surface mixing shoals to a depth shallower than a critical depth horizon defining the point where phytoplankton growth exceeds losses. This hypothesis has since served as a cornerstone in plankton ecology and re04ects the very common assumption that blooms are caused by enhanced growth rates in response to improved light, temperature, and strati03cation conditions, not simply correlated with them. Here, a nine-year satellite record of phytoplankton biomass in the subarctic Atlantic is used to reevaluate seasonal plankton dynamics. Results show that (1) bloom initiation occurs in the winter when mixed layer depths are maximum, not in the spring, (2) coupling between phytoplankton growth (l) and losses increases during spring strati03cation, rather than decreases, (3) maxima in net population growth rates (r) are as likely to occur in midwinter as in spring, and (4 ) r is generally inversely related to l. These results are incompatible with the Critical Depth Hypothesis as a functional framework for understanding bloom dynamics. In its place, a ‘‘Dilution–Recoupling Hypothesis’’ is described that focuses on the balance between phytoplankton growth and grazing, and the seasonally varying physical processes in04uencing this balance. This revised view derives from fundamental concepts applied during 03eld dilution experiments, builds upon earlier modeling results, and is compatible with observed phytoplankton blooms in the absence of spring mixed layer shoaling.
    Boss E., M. Behrenfeld, 2010: In situ evaluation of the initiation of the North Atlantic phytoplankton bloom. Geophys. Res. Lett., 37,L18603, doi: 10.1029/2010GL044174.10.1029/2010GL04417426cefdaa0e41dd0aff5300473d4ffe88http%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1029%2F2010GL044174%2Fcitedbyhttp://onlinelibrary.wiley.com/doi/10.1029/2010GL044174/citedbyTwo years of continuous physical and optical measurements from a profiling float in the western subarctic North Atlantic are used to analyze seasonal phytoplankton dynamics. The observed annual cycle challenges the traditional view that initiation of spring accumulations of phytoplankton in the upper water column requires a critical stratification threshold (known as the 'Gran effect' or the 'S...
    Chiswell S. M., 2011: Annual cycles and spring blooms in phytoplankton: Don't abandon Sverdrup completely. Marine Ecology Progress Series, 443, 39- 50.10.3354/meps094531ec4f077382522901e3bc56aa357dad4http%3A%2F%2Fwww.researchgate.net%2Fpublication%2F235289594_Annual_cycles_and_spring_blooms_in_phytoplankton_don%27t_abandon_Sverdrup_completelyhttp://www.researchgate.net/publication/235289594_Annual_cycles_and_spring_blooms_in_phytoplankton_don't_abandon_Sverdrup_completelyThe critical-depth model for the onset of the spring phytoplankton bloom in the North Atlantic has recently been called into question by several researchers. The critical-depth model considers that the spring bloom starts when the mixed layer shoals to become shallower than a critical depth. Satellite and in situ measurements of chlorophyll are used here to show that the critical-depth model is indeed flawed. It is shown that the critical-depth model does not apply in the spring because the basic assumption of an upper layer that is well-mixed in plankton is not met. Instead, the spring bloom forms in shallow near-surface layers that deepen with the onset of thermal stratification. A stratification-onset model for the annual cycle in plankton is proposed that adheres to the conventional idea that the spring bloom represents a change from a deep-mixed regime to a shallow light-driven regime, but where the upper layers are not well mixed in plankton in spring and so the critical-depth model does not apply. Ironically, perhaps, the critical-depth model applies in the autumn and winter when plankton are well-mixed to the seasonal thermocline, so that the critical-depth model can be used to determine whether net winter production is positive or negative.
    Chiswell S. M., P. H. R. Calil, and P. W. Boyd, 2015: Spring blooms and annual cycles of phytoplankton: a unified perspective. Journal of Plankton Research, 37( 3), 500- 508.10.1093/plankt/fbv021cad92d889b2ed45e2a4df516b015f279http%3A%2F%2Fplankt.oxfordjournals.org%2Fcontent%2F37%2F3%2F500.shorthttp://plankt.oxfordjournals.org/content/37/3/500.shortABSTRACT Several hypotheses exist that describe phytoplankton spring blooms in temperate and subpolar oceans: the critical depth, shoaling mixed layer (ML), critical turbulence, onset of stratification and disturbance-recovery hypotheses. These theories appear to be mutually exclusive and none of them describe the annual cycle of phytoplankton biomass. Here, we present a model of the annual cycle in phytoplankton that recognizes that phytoplankton are not always mixed throughout the so-called ML, and that it is important to distinguish between the surface biomass and depth-integrated phytoplankton. Once these important distinctions are made, the annual cycles and blooms in surface and depth-integrated phytoplankton can be described straightforwardly in terms of the physical drivers and biotic responses.
    Gill A. E., 1982: Atmosphere-Ocean Dynamics.Academic press, 662 pp.
    Hashioka T., T. T. Sakamoto, and Y. Yamanaka, 2009: Potential impact of global warming on north Pacific spring blooms projected by an eddy-permitting 3-D ocean ecosystem model. Geophys. Res. Lett., 36, L20604.10.1029/2009GL038912bb04c43a344b4804ae0e3a3f4a20b2b9http%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1029%2F2009GL038912%2Fcitedbyhttp://onlinelibrary.wiley.com/doi/10.1029/2009GL038912/citedbyUsing an eddy-permitting ecosystem model with a projected physical environment from a high-resolution climate model, we explored the potential impact of global warming on spring blooms in the western North Pacific. We focused on statistically significant signals compared with natural variability. Considering 2 CO2 conditions, maximum biomass during the spring bloom is found to occur 10 to 20 d...
    Held I. M., M. Winton, K. Takahashi, T. Delworth, F. R. Zeng, and G. K. Vallis, 2010: Probing the fast and slow components of global warming by returning abruptly to preindustrial forcing. J. Climate, 23( 9), 2418- 2427.10.1175/2009JCLI3466.1c4864a7f-71d6-4a41-ad26-9143e0e6047a724e0fd567acb7a9aa597d31a3d249b7http%3A%2F%2Fwww.cabdirect.org%2Fabstracts%2F20103193982.htmlrefpaperuri:(4e9122e2f4ebe1b6b817388f040a8cf1)http://www.cabdirect.org/abstracts/20103193982.htmlAbstract The fast and slow components of global warming in a comprehensive climate model are isolated by examining the response to an instantaneous return to preindustrial forcing. The response is characterized by an initial fast exponential decay with an e -folding time smaller than 5 yr, leaving behind a remnant that evolves more slowly. The slow component is estimated to be small at present, as measured by the global mean near-surface air temperature, and, in the model examined, grows to 0.4°C by 2100 in the A1B scenario from the Special Report on Emissions Scenarios (SRES), and then to 1.4°C by 2300 if one holds radiative forcing fixed after 2100. The dominance of the fast component at present is supported by examining the response to an instantaneous doubling of CO 2 and by the excellent fit to the model’s ensemble mean twentieth-century evolution with a simple one-box model with no long times scales.
    Huisman J. E. F., P. van Oostveen, and F. J. Weissing, 1999: Critical depth and critical turbulence: Two different mechanisms for the development of phytoplankton blooms. Limnology and Oceanography, 44( 7), 1781- 1787.10.4319/lo.1999.44.7.178157b6947e53b1d6a266c069e595423b10http%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.4319%2Flo.1999.44.7.1781%2Fpdfhttp://onlinelibrary.wiley.com/doi/10.4319/lo.1999.44.7.1781/pdfA turbulent diffusion model shows that there are two different mechanisms for the development of phytoplankton blooms. One of these mechanisms works in well-mixed environments and corresponds to the classical critical depth theory. The other mechanism is based on the rate of turbulent mixing. If turbulent mixing is less than a critical turbulence, phytoplankton growth rates exceed the vertical mixing rates, and a bloom develops irrespective of the depth of the upper water layer. These results demonstrate that phytoplankton blooms can develop in the absence of vertical water-column stratification.
    Kara A. B., P. A. Rochford, and H. E. Hurlburt, 2003: Mixed layer depth variability over the global ocean., J. Geophys. Res., 1083079, doi: 10.1029/2000C000736.10.1029/2000JC000736abd3c64f533899a42ab5fe0dccca1c14http%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1029%2F2000JC000736%2Ffullhttp://onlinelibrary.wiley.com/doi/10.1029/2000JC000736/fullAbstract Top of page Abstract 1.Introduction 2.Data and Limitations 3.Mixed Layer Processes and MLD Criterion 4.Overview of Global MLD Variability 5.ILD and MLD Correspondence 6.Validation of ILD Versus MLD Correspondence 7.Conclusions Acknowledgments References Supporting Information [1] The spatial and monthly variability of the climatological mixed layer depth (MLD) for the global ocean is examined using the recently developed Naval Research Laboratory (NRL) Ocean Mixed Layer Depth (NMLD) climatologies. The MLD fields are constructed using the subsurface temperature and salinity data from the World Ocean Atlas 1994 [ Levitus et al. , 1994 ; Levitus and Boyer , 1994 ]. To minimize the limitations of these global data in the MLD determination, a simple mixing scheme is introduced to form a stable water column. Using these new data sets, global MLD characteristics are produced on the basis of an optimal definition that employs a density-based criterion having a fixed temperature difference of T = 0.8C and variable salinity. Strong seasonality of MLD is found in the subtropical Pacific Ocean and at high latitudes, as well as a very deep mixed layer in the North Atlantic Ocean in winter and a very shallow mixed layer in the Antarctic in all months. Using the climatological monthly MLD and isothermal layer depth (ILD) fields from the NMLD climatologies, an annual mean T field is presented, providing criteria for determining an ILD that is approximately equivalent to the optimal MLD. This enables MLD to be determined in cases where salinity data are not available. The validity of the correspondence between ILD and MLD is demonstrated using daily averaged subsurface temperature and salinity from two moorings: a Tropical Atmosphere Ocean array mooring in the western equatorial Pacific warm pool, where salinity stratification is important, and a Woods Hole Oceanographic Institute (WHOI) mooring in the Arabian Sea, where strongly reversing seasonal monsoon winds prevail. In the western equatorial Pacific warm pool the use of ILD criterion with an annual mean T value of 0.3C yields comparable results with the optimal MLD, while large T values yield an overestimated MLD. An analysis of ILD and MLD in the WHOI mooring show that use of an incorrect T criterion for the ILD may underestimate or overestimate the optimal MLD. Finally, use of the spatial annual mean T values constructed from the NMLD climatologies can be used to estimate the optimal MLD from only subsurface temperature data via an equivalent ILD for any location over the global ocean.
    Kraus E. B., J. S. Turner, 1967: A one-dimensional model of the seasonal thermocline II. The general theory and its consequences. Tellus, 19( 1), 98- 106.10.1111/j.2153-3490.1967.tb01462.x56b73a1fa968e51c5a00ba71fef2fc70http%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1111%2Fj.2153-3490.1967.tb01462.x%2Fcitedbyhttp://onlinelibrary.wiley.com/doi/10.1111/j.2153-3490.1967.tb01462.x/citedbyA theory of the layer formation due to surface processes is presented, which is more general than that used in the preceding paper I. Convection due to heating at depth and cooling at the surface is included, as well as the mechanical stirring due to wind action. The theory is applicable to arbitrary forms of heating, including intermittent or continuous processes, and could be used to investigate diurnal as well as seasonal effects. A detailed application is made to the case treated approximately in I, for which a solution is now obtained in analytic form.The results obtained allow a quantitative, as well as qualitative, comparison with the ocean. It is found that reasonable layer depths are predicted using measured heating rates, and a value of the turbulent kinetic energy input to the water deduced from the mean surface stress. The effects of heating at depth can be comparable with wind stirring, even when the temperature of the upper layer is increasing. During the winter, convection due to surface cooling dominates the processes which deepen the layer.
    Kraus E. B., J. A. Businger, 1995: Atmosphere-Ocean Interaction,2nd ed., Oxford University Press, 362 pp.10.1017/S002211207321162Xbc1db51819a36070389721e11b6e9802http%3A%2F%2Fci.nii.ac.jp%2Fncid%2FBA23771309http://ci.nii.ac.jp/ncid/BA23771309The representative of Department of investigation of flight accidents of the Ministry of transport of the great Britain has informed, that on the facts of the crash will be investigated. In sezone-2009 10 became the best assistant among the defenders in the Finnish
    Long S.-M., S.-P. Xie, X.-T. Zheng, and Q. Y. Liu, 2014: Fast and slow responses to global warming: Sea surface temperature and precipitation patterns. J. Climate, 27( 1), 285- 299.10.1175/JCLI-D-13-00297.15aef149cc91f2d5d11bce02b01ae1617http%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F2014JCli...27..285Lhttp://adsabs.harvard.edu/abs/2014JCli...27..285LThe time-dependent response of sea surface temperature (SST) to global warming and the associated atmospheric changes are investigated based on a 1% yr 1 CO2 increase to the quadrupling experiment of the Geophysical Fluid Dynamics Laboratory Climate Model, version 2.1. The SST response consists of a fast component, for which the ocean mixed layer is in quasi equilibrium with the radiative forcing, and a slow component owing to the gradual warming of the deeper ocean in and beneath the thermocline. A diagnostic method is proposed to isolate spatial patterns of the fast and slow responses. The deep ocean warming retards the surface warming in the fast response but turns into a forcing for the slow response. As a result, the fast and slow responses are nearly opposite to each other in spatial pattern, especially over the subpolar North Atlantic/Southern Ocean regions of the deep-water/bottom-water formation, and in the interhemispheric SST gradient between the southern and northern subtropics. Wind-evaporation-SST feedback is an additional mechanism for the SST pattern formation in the tropics. Analyses of phase 5 of the Coupled Model Intercomparison Project (CMIP5) multimodel ensemble of global warming simulations confirm the validity of the diagnostic method that separates the fast and slow responses. Tropical annual rainfall change follows the SST warming pattern in both the fast and slow responses in CMIP5, increasing where the SST increase exceeds the tropical mean warming.
    Luo Y. Y., Q. Y. Liu, and L. M. Rothstein, 2009: Simulated response of north Pacific mode waters to global warming. Geophys. Res. Lett., 36,L23609, doi: 10.1029/2009GL040906.10.1029/2009GL040906263773f58656265a816fc286db80caa2http%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1029%2F2009GL040906%2Fabstracthttp://onlinelibrary.wiley.com/doi/10.1029/2009GL040906/abstract[1] This study investigates the response of the Mode Waters in the North Pacific to global warming based on a set of Intergovernmental Panel on Climate Change (IPCC) Fourth Assessment Report (AR4) models. Solutions between a present-day climate and a future, warmer climate are compared. Under the warmer climate scenario, the Mode Waters are produced on lighter isopycnal surfaces and are significantly weakened in terms of their formation and evolution. These changes are due to a more stratified upper ocean and thus a shoaling of the winter mixing depth resulting mainly from a reduction of the ocean-to-atmosphere heat loss over the subtropical region. The basin-wide wind stress may adjust the Mode Waters indirectly through its impact on the surface heat flux and subduction process.
    Mann K., J. Lazier, 2005: Dynamics of Marine Ecosystems: Biological-Physical Interactions in the Oceans. Wiley-Blackwell,496 pp.10.1016/S0025-326X(97)00072-613fa9ec7385291288c1040b2020861fbhttp%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1111%2Fj.1365-2419.1992.tb00029.x%2Fabstracthttp://onlinelibrary.wiley.com/doi/10.1111/j.1365-2419.1992.tb00029.x/abstractNo abstract is available for this article.
    Murtugudde R., J. Beauchamp, C. R. McClain, M. Lewis, and A. J. Busalacchi, 2002: Effects of penetrative radiation on the upper tropical Ocean circulation. J. Climate, 15( 5), 470- 486.10.1175/1520-0442(2002)015<0470:EOPROT>2.0.CO;285d29fd84f8b723106ec13fdd16a27e7http%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F2002JCli...15..470Mhttp://adsabs.harvard.edu/abs/2002JCli...15..470MThe effects of penetrative radiation on the upper tropical ocean circulation have been investigated with an ocean general circulation model (OGCM) with attenuation depths derived from remotely sensed ocean color data. The OGCM is a reduced gravity, primitive equation, sigma coordinate model coupled to an advective atmospheric mixed layer model. These simulations use a single exponential profile for radiation attenuation in the water column, which is quite accurate for OGCMs with fairly coarse vertical resolution. The control runs use an attenuation depth of 17 m while the simulations use spatially variable attenuation depths. When a variable depth oceanic mixed layer is explicitly represented with interactive surface heat fluxes, the results can be counterintuitive. In the eastern equatorial Pacific, a tropical ocean region with one of the strongest biological activity, the realistic attenuation depths result in increased loss of radiation to the subsurface, but result in increased sea surface temperatures (SSTs) compared to the control run. Enhanced subsurface heating leads to weaker stratification, deeper mixed layers, reduced surface divergence, and hence less upwelling and entrainment. Thus, some of the systematic deficiencies in the present-day climate models, such as the colder than observed cold tongue in the equatorial Pacific may simply be related to inaccurate representation of the penetrative radiation and can be improved by the formulation presented here. The differences in ecosystems in each of the tropical oceans are clearly manifested in the manner in which biological heat trapping affects the upper ocean. While the tropical Atlantic has many similarities to the Pacific, the Amazon, Congo, and Niger Rivers' discharges dominate the attenuation of radiation. In the Indian Ocean, elevated biological activity and heat trapping are away from the equator in the Arabian Sea and the southern Tropics. For climate models, in view of their sensitivity to...
    Nakamoto S., S. P. Kumar, J. M. Oberhuber, J. Ishizaka, K. Muneyama, and R. Frouin, 2001: Response of the equatorial Pacific to chlorophyll pigment in a mixed layer isopycnal ocean general circulation model. Geophys. Res. Lett., 28( 10), 2021- 2024.10.1029/2000GL01249489ef5d21154a53bd9abb70e209efbe21http%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1029%2F2000GL012494%2Fabstracthttp://onlinelibrary.wiley.com/doi/10.1029/2000GL012494/abstractThe influence of phytoplankton on the upper ocean dynamics and thermodynamics in the equatorial Pacific is investigated using an isopycnal ocean general circulation model (OPYC) coupled with a mixed layer model and remotely sensed chlorophyll pigment concentration from the Coastal Zone Color Scanner (CZCS). In the equatorial Pacific heat accumulation due to a higher abundance of chlorophyll pigments in the equatorial Pacific leads to a decrease of the mixed layer thickness. This generates anomalous westward geostrophic currents north and south of the equator. In the western equatorial Pacific, these anomalous geostrophic currents merge into and strengthen the equatorial undercurrent (EUC), supplying water mass from the 200 m depth to the eastern equatorial Pacific. This chlorophyll-induced response of the undercurrent enhances upwelling around 110W, resulting in a lower sea surface temperature (SST) than without chlorophyll. Thus, thermal gradients due to absorption of solar radiation by phytoplankton may contribute remotely to equatorial upwelling in the eastern Pacific.
    Qiu B., K. A. Kelly, 1993: Upper-ocean heat balance in the Kuroshio extension region. J. Phys. Oceanogr., 23( 9), 2027- 2041.10.1175/1520-0485(1993)023<2027:UOHBIT>2.0.CO;2bc27bdcf4fbd2a076bd1e672aac8b38ahttp%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F1993JPO....23.2027Qhttp://adsabs.harvard.edu/abs/1993JPO....23.2027QAbstract A horizontally two-dimensional mixed-layer model is used to study the upper-ocean heat balance in the Kuroshio Extension region (30°–40°N, 141°–175°E). Horizontal dependency is emphasized because, in addition to vertical entrainment and surface thermal forcing, horizontal advection and eddy diffusion make substantial contributions to changes in the upper-ocean thermal structure in this region. By forcing the model using the wind and heat flux data from ECMWF and the absolute sea surface height data deduced from the Geosat ERM, the mixed-layer depth ( h m ) and temperature ( T m ) changes in the Kuroshio Extension are hindcast for a 2.5-year period (November 1986–April 1989). Both phase and amplitude of the modeled T m and h m variations agreed well with the climatology. The horizontal thermal patterns also agreed favorably with the available in situ SST observations, but this agreement depended crucially on the inclusion of horizontal advections. Although the annually averaged net heat flux from the atmosphere to the ocean ( Q net ) is negative over the Kuroshio Extension region, the effect of the surface thermal forcing, when integrated annually, is to increase T m because the large, negative Q net in winter is redistributed in a much deeper mixed layer than it is in summer when Q net > 0. This warming effect is counterbalanced by the vertical turbulent entrainment through the base of the mixed layer (35% when annually integrated), the Ekman divergence (16%), the geostrophic divergence (12%), and the horizontal eddy diffusion (35%). Though small when averaged in space and time, the temperature advection by the surface flows makes a substantial contribution to the local heat balances. While it warms the upstream region of the Kuroshio Extension (west of 150°E), the current advection tends to cool the upper ocean over the vast downstream region due to the presence of the recirculation gyre.
    Qiu B., N. Schneider, and S. M. Chen, 2007: Coupled decadal variability in the north pacific: An observationally constrained idealized model. J. Climate, 20( 14), 3602- 3620.a9fb9946-e5b6-47ad-a2cb-f71ea0884c93a213cd08b419c39ba72e5ee5e9aa6221http%3A%2F%2Fwww.nrcresearchpress.com%2Fservlet%2Flinkout%3Fsuffix%3Drg20%2Fref20%26dbid%3D16%26doi%3D10.1139%252FF09-051%26key%3D10.1175%252FJCLI4190.1refpaperuri:(a2fa8f28e90550e6176a8f216a50822a)/s?wd=paperuri%3A%28a2fa8f28e90550e6176a8f216a50822a%29&filter=sc_long_sign&tn=SE_xueshusource_2kduw22v&sc_vurl=http%3A%2F%2Fwww.nrcresearchpress.com%2Fservlet%2Flinkout%3Fsuffix%3Drg20%2Fref20%26dbid%3D16%26doi%3D10.1139%252FF09-051%26key%3D10.1175%252FJCLI4190.1&ie=utf-8
    Sakamoto T. T., H. Hasumi, M. Ishii, S. Emori, T. Suzuki, T. Nishimura, and A. Sumi, 2005: Responses of the Kuroshio and the Kuroshio extension to global warming in a high-resolution climate model. Geophys. Res. Lett., 32,L14617, doi: 10.1029/2005GL023384.10.1029/2005GL0233840c5824bdbd736629b636fdd6b939d90chttp%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1029%2F2005GL023384%2Ffullhttp://onlinelibrary.wiley.com/doi/10.1029/2005GL023384/fullABSTRACT Using a high-resolution atmosphere-ocean coupled climate model, responses of the Kuroshio and the Kuroshio Extension (KE) to global warming are investigated. In a climate change experiment with atmospheric CO2 concentration ideally increased by 1% year-1, the current velocity of the Kuroshio and KE increases, while the latitude of the Kuroshio separation to the east of Japan does not change significantly. The increase of the current velocity is up to 0.3 m s-1 at 150E. This acceleration of the Kuroshio and KE is due to changes in wind stress over the North Pacific and consequent spin-up of the Kuroshio recirculation gyre. The acceleration of the currents may affect sea level along the southern coast of Japan and northward heat transport under global warming.
    Sato Y., S. Yukimoto, H. Tsujino, H. Ishizaki, and A. Noda, 2006: Response of North Pacific Ocean circulation in a Kuroshio-resolving ocean model to an Arctic Oscillation (AO)-like change in Northern Hemisphere atmospheric circulation due to greenhouse-gas forcing. J. Meteor. Soc.Japan, 84, 295- 309.10.2151/jmsj.84.29585edcbbf-9867-4e4a-aecb-2af2235058f467128fb67d2857446cd3667232e169fahttp%3A%2F%2Fci.nii.ac.jp%2Fnaid%2F130004434931refpaperuri:(a4e814ef6450a3b2d98e4644ee767897)http://ci.nii.ac.jp/naid/130004434931Time-slice experiments are performed using a high-resolution North Pacific ocean general circulation model (NPOGCM) resolving the strong currents near Japan, such as the Kuroshio and the Oyashio, to investigate the effect of global warming on the North Pacific ocean circulation. The NPOGCM is forced by heat, momentum, and fresh-water fluxes obtained from a global warming projection using a global climate model (MRI-CGCM2.2). The annual mean sea-level pressure trend exhibits an annular pattern similar to the positive phase of the Arctic Oscillation in a global warming projection by MRI-CGCM2.2 based on the Intergovernmental Panel on Climate Change (IPCC) SRES A2 emission scenario. Associated with this trend, the anticyclonic atmospheric circulation is intensified over the mid-latitude North-Pacific, leading to a north-ward shift of the oceanic subtropical wind-driven gyre boundary, where extensions of the Kuroshio exist in MRI-CGCM2.2. Under these forcing changes, NPOGCM projects that in the future climate warm core eddies are more frequently pinched off from the Kuroshio off the eastern coast of Japan, leading to an annual mean SST rise over 5K at its maximum, compared with the present climate. The projected annual mean sea-level rise ranges from 12 to 18 cm along the coasts of Japan, and about 40 cm over the ocean east of Japan.
    Sverdrup H. U., 1953: On conditions for the vernal blooming of phytoplankton. Journal du Conseil International Pour l'Exploration de la Mer, 18( 3), 287- 295.10.1093/icesjms/18.3.287152c4ceec2141faba88c5a1a27a3277ehttp%3A%2F%2Ficesjms.oxfordjournals.org%2Fcontent%2F18%2F3%2F287.full.pdfhttp://icesjms.oxfordjournals.org/content/18/3/287.full.pdfPublication &raquo; On conditions for the vernal blooming of phytoplankton.
    Taguchi B., S.-P. Xie, N. Schneider, M. Nonaka, H. Sasaki, and Y. Sasai, 2007: Decadal variability of the Kuroshio extension: observations and an eddy-resolving model hindcast. J. Climate, 20( 11), 2357- 2377.10.1175/JCLI4142.172b9f343a411bb80b3e3bc88f519aed2http%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F2007JCli...20.2357Thttp://adsabs.harvard.edu/abs/2007JCli...20.2357TLow-frequency variability of the Kuroshio Extension (KE) is studied using observations and a multidecadal (195009-2003) hindcast by a high-resolution (0.100°), eddy-resolving, global ocean general circulation model for the Earth Simulator (OFES). In both the OFES hindcast and satellite altimeter observations, low-frequency sea surface height (SSH) variability in the North Pacific is high near the KE front. An empirical orthogonal function (EOF) analysis indicates that much of the SSH variability in the western North Pacific east of Japan is explained by two modes with meridional structures tightly trapped along the KE front. The first mode represents a southward shift and to a lesser degree, an acceleration of the KE jet associated with the 1976/77 shift in basin-scale winds. The second mode reflects quasi-decadal variations in the intensity of the KE jet. Both the spatial structure and time series of these modes derived from the hindcast are in close agreement with observations. A linear Rossby wave model forced by observed wind successfully reproduces the time series of the leading OFES modes but fails to explain why their meridional structure is concentrated on the KE front and inconsistent with the broadscale wind forcing. Further analysis suggests that KE variability may be decomposed into broad- and frontal-scale components in the meridional direction09恪眛he former following the linear Rossby wave solution and the latter closely resembling ocean intrinsic modes derived from an OFES run forced by climatological winds. The following scenario is suggested for low-frequency KE variability: basin-scale wind variability excites broadscale Rossby waves, which propagate westward, triggering intrinsic modes of the KE jet and reorganizing SSH variability in space.
    Taylor J. R., R. Ferrari, 2011: Shutdown of turbulent convection as a new criterion for the onset of spring phytoplankton blooms. Limnology and Oceanography, 56( 6), 2293- 2307.10.4319/lo.2011.56.6.22937b895f1c-672b-4824-9e48-1c6ee79abfc08cb8f0b7cd9f0ab1ad5d934cef8e52d9http%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.4319%2Flo.2011.56.6.2293%2Fcitedbyrefpaperuri:(ee4c72d26d5e74505553a3ff0eccce56)http://onlinelibrary.wiley.com/doi/10.4319/lo.2011.56.6.2293/citedbyThe onset of phytoplankton blooms in late winter, early spring has been traditionally associated with the shoaling of the mixed layer above a critical depth. Here we show that the onset of a bloom can also be triggered by a reduction in air–sea fluxes at the end of winter. When net cooling subsides at the end of winter, turbulent mixing becomes weak, thereby increasing the residence time of phytoplankton cells in the euphotic layer and allowing a bloom to develop. The necessary change in the air–sea flux generally precedes mixed-layer shoaling, and may provide a better indicator for the onset of the spring bloom than the mixed-layer depth alone. Our hypothesis is supported by numerical simulations and remote sensing data.
    Taylor K. E., R. J. Stouffer, and G. A. Meehl, 2012: An overview of CMIP5 and the experiment design. Bull. Amer. Meteor. Soc., 93( 4), 485- 498.10.1175/BAMS-D-11-00094.10a93ff62-7ac1-4eaa-951b-da834bb5d6acd378bae55de68ca8b37ba4ba57a3c0b9http%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F2012BAMS...93..485Trefpaperuri:(102c64f576f0dc49ca552e6df691421b)http://adsabs.harvard.edu/abs/2012BAMS...93..485TThe fifth phase of the Coupled Model Intercomparison Project (CMIP5) will produce a state-of-the- art multimodel dataset designed to advance our knowledge of climate variability and climate change. Researchers worldwide are analyzing the model output and will produce results likely to underlie the forthcoming Fifth Assessment Report by the Intergovernmental Panel on Climate Change. Unprecedented in scale and attracting interest from all major climate modeling groups, CMIP5 includes “long term” simulations of twentieth-century climate and projections for the twenty-first century and beyond. Conventional atmosphere–ocean global climate models and Earth system models of intermediate complexity are for the first time being joined by more recently developed Earth system models under an experiment design that allows both types of models to be compared to observations on an equal footing. Besides the longterm experiments, CMIP5 calls for an entirely new suite of “near term” simulations focusing on recent decades and the future to year 2035. These “decadal predictions” are initialized based on observations and will be used to explore the predictability of climate and to assess the forecast system's predictive skill. The CMIP5 experiment design also allows for participation of stand-alone atmospheric models and includes a variety of idealized experiments that will improve understanding of the range of model responses found in the more complex and realistic simulations. An exceptionally comprehensive set of model output is being collected and made freely available to researchers through an integrated but distributed data archive. For researchers unfamiliar with climate models, the limitations of the models and experiment design are described.
    Vecchi G. A., B. J. Soden, 2007: Global warming and the weakening of the tropical circulation. J. Climate, 20( 17), 4316- 4340.10.1175/JCLI4258.1a806368e-bebb-4d5b-ad72-296f452686b473c4454bd5c87d67e3bd17752bb3e9d7http%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F2006AGUFMOS11A1467Wrefpaperuri:(b30ccd0a9e8c71855c1ae5c8c7674dc2)http://adsabs.harvard.edu/abs/2006AGUFMOS11A1467WAbstract This study examines the response of the tropical atmospheric and oceanic circulation to increasing greenhouse gases using a coordinated set of twenty-first-century climate model experiments performed for the Intergovernmental Panel on Climate Change (IPCC) Fourth Assessment Report (AR4). The strength of the atmospheric overturning circulation decreases as the climate warms in all IPCC AR4 models, in a manner consistent with the thermodynamic scaling arguments of Held and Soden. The weakening occurs preferentially in the zonally asymmetric (i.e., Walker) rather than zonal-mean (i.e., Hadley) component of the tropical circulation and is shown to induce substantial changes to the thermal structure and circulation of the tropical oceans. Evidence suggests that the overall circulation weakens by decreasing the frequency of strong updrafts and increasing the frequency of weak updrafts, although the robustness of this behavior across all models cannot be confirmed because of the lack of data. As the climate warms, changes in both the atmospheric and ocean circulation over the tropical Pacific Ocean resemble “El Ni09o–like” conditions; however, the mechanisms are shown to be distinct from those of El Ni09o and are reproduced in both mixed layer and full ocean dynamics coupled climate models. The character of the Indian Ocean response to global warming resembles that of Indian Ocean dipole mode events. The consensus of model results presented here is also consistent with recently detected changes in sea level pressure since the mid–nineteenth century.
    Wu L., Z. Liu, R. Gallimore, R. Jacob, D. Lee, and Y. Zhong, 2003: Pacific decadal variability: The tropical Pacific mode and the north Pacific mode. J. Climate, 16( 8), 1101- 1120.21717738-dafd-4061-ad47-3c28e28c76d15a12d010ade3947e5b1cbf4adcbf62f8http%3A%2F%2Fadsabs.harvard.edu%2Fcgi-bin%2Fnph-data_query%3Fbibcode%3D2003JCli...16.1101W%26db_key%3DPHY%26link_type%3DABSTRACT%26high%3D02861refpaperuri:(983b2f921f0aca1092e170b0258c11c9)/s?wd=paperuri%3A%28983b2f921f0aca1092e170b0258c11c9%29&filter=sc_long_sign&tn=SE_xueshusource_2kduw22v&sc_vurl=http%3A%2F%2Fadsabs.harvard.edu%2Fcgi-bin%2Fnph-data_query%3Fbibcode%3D2003JCli...16.1101W%26db_key%3DPHY%26link_type%3DABSTRACT%26high%3D02861&ie=utf-8
    Wu, L. X., Coauthors , 2012: Enhanced warming over the global subtropical western boundary currents. Nature Climate Change, 2( 3), 161- 166.10.1038/nclimate1353c598309ed8e934930faec5c8b0219d66http%3A%2F%2Fwww.nature.com%2Fnclimate%2Fjournal%2Fv2%2Fn3%2Fnclimate1353%2Fmetricshttp://www.nature.com/nclimate/journal/v2/n3/nclimate1353/metricsSubtropical western boundary currents are warm, fast-flowing currents that form on the western side of ocean basins. They carry warm tropical water to the mid-latitudes and vent large amounts of heat and moisture to the atmosphere along their paths, affecting atmospheric jet streams and mid-latitude storms, as well as ocean carbon uptake. The possibility that these highly energetic currents might change under greenhouse-gas forcing has raised significant concerns, but detecting such changes is challenging owing to limited observations. Here, using reconstructed sea surface temperature datasets and century-long ocean and atmosphere reanalysis products, we find that the post-1900 surface ocean warming rate over the path of these currents is two to three times faster than the global mean surface ocean warming rate. The accelerated warming is associated with a synchronous poleward shift and/or intensification of global subtropical western boundary currents in conjunction with a systematic change in winds over both hemispheres. This enhanced warming may reduce the ability of the oceans to absorb anthropogenic carbon dioxide over these regions. However, uncertainties in detection and attribution of these warming trends remain, pointing to a need for a long-term monitoring network of the global western boundary currents and their extensions.
    Xie S.-P., C. Deser, G. A. Vecchi, J. Ma, H. Y. Teng, and A. T. Wittenberg, 2010: Global warming pattern formation: Sea surface temperature and rainfall. J. Climate, 23( 4), 966- 986.10.1175/2009JCLI3329.13eff0181-4d63-488f-92b9-71a15a93bf303eefc59c87e4f4ca7ffcbe050a36a436http%3A%2F%2Fwww.cabdirect.org%2Fabstracts%2F20103118162.htmlrefpaperuri:(b0ebeb07b4f54809d624dfe9936fb36c)http://www.cabdirect.org/abstracts/20103118162.htmlAbstract Spatial variations in sea surface temperature (SST) and rainfall changes over the tropics are investigated based on ensemble simulations for the first half of the twenty-first century under the greenhouse gas (GHG) emission scenario A1B with coupled ocean tmosphere general circulation models of the Geophysical Fluid Dynamics Laboratory (GFDL) and National Center for Atmospheric Research (NCAR). Despite a GHG increase that is nearly uniform in space, pronounced patterns emerge in both SST and precipitation. Regional differences in SST warming can be as large as the tropical-mean warming. Specifically, the tropical Pacific warming features a conspicuous maximum along the equator and a minimum in the southeast subtropics. The former is associated with westerly wind anomalies whereas the latter is linked to intensified southeast trade winds, suggestive of wind vaporation ST feedback. There is a tendency for a greater warming in the northern subtropics than in the southern subtropics in accordance with asymmetries in trade wind changes. Over the equatorial Indian Ocean, surface wind anomalies are easterly, the thermocline shoals, and the warming is reduced in the east, indicative of Bjerknes feedback. In the midlatitudes, ocean circulation changes generate narrow banded structures in SST warming. The warming is negatively correlated with wind speed change over the tropics and positively correlated with ocean heat transport change in the northern extratropics. A diagnostic method based on the ocean mixed layer heat budget is developed to investigate mechanisms for SST pattern formation. Tropical precipitation changes are positively correlated with spatial deviations of SST warming from the tropical mean. In particular, the equatorial maximum in SST warming over the Pacific anchors a band of pronounced rainfall increase. The gross moist instability follows closely relative SST change as equatorial wave adjustments flatten upper-tropospheric warming. The comparison with atmospheric simulations in response to a spatially uniform SST warming illustrates the importance of SST patterns for rainfall change, an effect overlooked in current discussion of precipitation response to global warming. Implications for the global and regional response of tropical cyclones are discussed.
    Xu L. X., S.-P. Xie, and Q. Y. Liu, 2012: Mode water ventilation and subtropical countercurrent over the north Pacific in CMIP5 simulations and future projections. J. Geophys. Res.-Oceans, 117,C12009, doi: 10.1029/2012JC008377.10.1029/2012JC00837722752837e64d9cfa-d34e-4397-8b18-9c854ab6cdc571738c2fd07a32ad54476b4634641199http%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1029%2F2012JC008377%2Fcitedbyrefpaperuri:(a457eb87eca80cea5de2eddefbe2c0d4)http://onlinelibrary.wiley.com/doi/10.1029/2012JC008377/citedby[1] Seventeen coupled general circulation models from the Coupled Model Intercomparison Project Phase 5 (CMIP5) are analyzed to assess the dynamics and variability of the North Pacific Subtropical Countercurrent (STCC). Consistent with observations, the STCC is anchored by mode water to the north. For the present climate, the STCC tends to be stronger in models than in observations because of too strong a low potential vorticity signature of mode water. There are significant variations in mode water simulation among models, i.e., in volume and core layer density. The northeast slanted bands of sea surface height (SSH) anomalies associated with the STCC variability are caused by variability in mode water among models and the Hawaii islands are represented in some models, where the island-induced wind curls drive the Hawaiian Lee Countercurrent (HLCC) located to the south of STCC. Projected future changes in STCC and mode water under the Representative Concentration Pathways (RCP) 4.5 scenario are also investigated. By combining the historical and RCP 4.5 runs, an empirical orthogonal function analysis for SSH over the central subtropical gyre (160 E&ndash;140W, 15 &ndash;30N) is performed. The dominant mode of SSH change in 17 CMIP5 models is characterized by the weakening of the STCC because of the reduced formation of mode water. The weakened mode water is closely related to the increased stratification of the upper ocean, the latter being one of the most robust changes as climate warms. Thus the weakened STCC and mode water are common to CMIP5 future climate projections.
    Xu L. X., S.-P. Xie, and Q. Y. Liu, 2013: Fast and slow responses of the north Pacific mode water and subtropical countercurrent to global warming. Journal of Ocean University of China, 12( 2), 216- 221.10.1007/s11802-013-2189-6ef113ecd-21ab-47b1-a736-8bc0ee2f73caWOS:000318300200004Six coupled general circulation models from the Coupled Model Intercomparison Project Phase 5 (CMIP5) are employed for examining the full evolution of the North Pacific mode water and Subtropical Countercurrent (STCC) under global warming over 400 years following the Representative Concentration Pathways (RCP) 4.5. The mode water and STCC first show a sharp weakening trend when the radiative forcing increases, but then reverse to a slow strengthening trend of smaller magnitude after the radiative forcing is stablized. As the radiative forcing increases during the 21st century, the ocean warming is surface-intensified and decreases with depth, strengthening the upper ocean's stratification and becoming unfavorable for the mode water formation. Moving southward in the subtropical gyre, the shrinking mode water decelerates the STCC to the south. After the radiative forcing is stabilized in the 2070s, the subsequent warming is greater at the subsurface than at the sea surface, destabilizing the upper ocean and becoming favorable for the mode water formation. As a result, the mode water and STCC recover gradually after the radiative forcing is stabilized.
    Xu L. X., S.-P. Xie, J. L. McClean, Q. Y. Liu, and H. Sasaki, 2014: Mesoscale eddy effects on the subduction of north Pacific mode waters. J. Geophys. Res.,119, 4867-4886, doi: 10.1002/2014JC009861.10.1002/2014JC009861c36cb47bf8cc6fc905ae1856eba81c45http%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1002%2F2014JC009861%2Fabstracthttp://onlinelibrary.wiley.com/doi/10.1002/2014JC009861/abstractAbstract Mesoscale eddy effects on the subduction of North Pacific mode waters are investigated by comparing observations and ocean general circulation models where eddies are either parameterized or resolved. The eddy-resolving models produce results closer to observations than the noneddy-resolving model. There are large discrepancies in subduction patterns between eddy-resolving and noneddy-resolving models. In the noneddy-resolving model, subduction on a given isopycnal is limited to the cross point between the mixed layer depth (MLD) front and the outcrop line whereas in eddy-resolving models and observations, subduction takes place in a broader, zonally elongated band within the deep mixed layer region. Mesoscale eddies significantly enhance the total subduction rate, helping create remarkable peaks in the volume histogram that correspond to North Pacific subtropical mode water (STMW) and central mode water (CMW). Eddy-enhanced subduction preferentially occurs south of the winter mean outcrop. With an anticyclonic eddy to the west and a cyclonic eddy to the east, the outcrop line meanders south, and the thermocline/MLD shoals eastward. As eddies propagate westward, the MLD shoals, shielding the water of low potential vorticity from the atmosphere. The southward eddy flow then carries the subducted water mass into the thermocline. The eddy subduction processes revealed here have important implications for designing field observations and improving models.
    Yentsch C. S., 1990: Estimates of `new production' in the Mid-North Atlantic. Journal of Plankton Research, 12, 717- 734.10.1093/plankt/12.4.7177392cb56-3d2d-4197-ac9f-659933ee9e6a344a42d33c8fbadb0b1af3c9dda2df02http%3A%2F%2Fplankt.oxfordjournals.org%2Fcontent%2F12%2F4%2F717.shortrefpaperuri:(ae2af3ec447c817bbac8afc2a792751c)http://plankt.oxfordjournals.org/content/12/4/717.shortAbstract The principal aim of this paper is to demonstrate how the major features of primary production are influenced by climatological change. The roles played by seasonal change in mixed layer depth and the vertical distribution of a key limiting nutrient are emphasized. The model traces the sequences of primary production as the sun moves from the winter to summer solstice. Seasonal change in primary production is regulated by light at high latitudes during winter months and nutrients during summer months, where at mid and low latitudes nutrients limit production. The annual pattern of production down the central meridian reflects the vertical distribution of nitrate-nitrogen in cross-sections of the mid-ocean. In turn, this pattern of nutrient distribution reflects the density structure due to the currents and gyres of the North Atlantic. The model produces estimates of ‘new primary production’ which are consistent when compared with measured values. It should be useful for global estimates of primary production.
    Yim B. Y., Y. Noh, S. W. Yeh, J. S. Kug, H. S. Min, and B. Qiu, 2013: Ocean mixed layer processes in the Pacific decadal oscillation in coupled general circulation models. Climate Dyn., 41( 5-6), 1407- 1417.10.1007/s00382-012-1630-7cf2953b785bbc468a22f79ac616f6a4chttp%3A%2F%2Flink.springer.com%2F10.1007%2Fs00382-012-1630-7http://link.springer.com/10.1007/s00382-012-1630-7It is investigated how the Pacific Decadal Oscillation (PDO) is simulated differently among various coupled general circulation models (CGCMs), and how it is related to the heat budget of the simulated ocean mixed layer, which includes the surface heat flux and ocean heat transport. For this purpose the dataset of the climate of the 20th Century experiment (20C3M) from nine CGCMs reported to IPCC AR4 are used, while the MRI and MIROC models are examined in detail. Detailed analyses of these two CGCMs reveal that the PDO is mainly affected by ocean heat transport rather than surface heat flux, in particular in the MRI model which has a larger contribution of ocean heat transport to the heat budget. It is found that the ocean heat transport due to Ekman advection versus geostrophic advection contributes differently to the PDO in the western and central North Pacific. Specifically, the strength of PDO tends to be larger for CGCMs with a larger ocean heat transport in the region.
    Yu L., X. Jin, and R. A. Weller, 2006: Role of net surface heat flux in seasonal variations of sea surface temperature in the tropical Atlantic Ocean. J. Climate, 19( 23), 6153- 6169.10.1175/JCLI3970.184bca3739e4f166686cbb3fcc1c81ea9http%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F2006JCli...19.6153Yhttp://adsabs.harvard.edu/abs/2006JCli...19.6153YThe present study used a new net surface heat flux (Qnet) product obtained from the Objective Analyzed Air09 ea Fluxes (OAFlux) project and the International Satellite Cloud Climatology Project (ISCCP) to examine two specific issues09ne is to which degree Qnet controls seasonal variations of sea surface temperature (SST) in the tropical Atlantic Ocean (2000°S09 2000°N, east of 6000°W), and the other is whether the physical relation can serve as a measure to evaluate the physical representation of a heat flux product. To better address the two issues, the study included the analysis of three additional heat flux products: the Southampton Oceanographic Centre (SOC) heat flux analysis based on ship reports, and the model fluxes from the National Centers for Environmental Prediction09 ational Center for Atmospheric Research (NCEP09 CAR) reanalysis and the 40-yr European Centre for Medium-Range Weather Forecasts (ECMWF) Re-Analysis (ERA-40). The study also uses the monthly subsurface temperature fields from the World Ocean Atlas to help analyze the seasonal changes of the mixed layer depth (hMLD). The study showed that the tropical Atlantic sector could be divided into two regimes based on the influence level of Qnet. SST variability poleward of 500°S and 1000°N is dominated by the annual cycle of Qnet. In these regions the warming (cooling) of the sea surface is highly correlated with the increased (decreased) Qnet confined in a relatively shallow (deep) hMLD. The seasonal evolution of SST variability is well predicted by simply relating the local Qnet with a variable hMLD. On the other hand, the influence of Qnet diminishes in the deep Tropics within 500°S and 1000°N and ocean dynamic processes play a dominant role. The dynamics-induced changes in SST are most evident along the two belts, one of which is located on the equator and the other off the equator at about 300°N in the west, which tilts to about 1000°N near the northwestern African coast. The study also showed that if the degree of consistency between the correlation relationships of Qnet, hMLD, and SST variability serves as a measure of the quality of the Qnet product, then the Qnet from OAFlux + ISCCP and ERA-40 are most physically representative, followed by SOC. The NCEP09 CAR Qnet is least representative. It should be noted that the Qnet from OAFlux + ISCCP and ERA-40 have a quite different annual mean pattern. OAFlux + ISCCP agrees with SOC in that the tropical Atlantic sector gains heat from the atmosphere on the annual mean basis, where the ERA-40 and the NCEP09 CAR model reanalyses indicate that positive Qnet occurs only in the narrow equatorial band and in the eastern portion of the tropical basin. Nevertheless, seasonal variances of the Qnet from OAFlux + ISCCP and ERA-40 are very similar once the respective mean is removed, which explains why the two agree with each other in accounting for the seasonal variability of SST. In summary, the study suggests that an accurate estimation of surface heat flux is crucially important for understanding and predicting SST fluctuations in the tropical Atlantic Ocean. It also suggests that future emphasis on improving the surface heat flux estimation should be placed more on reducing the mean bias.
  • [1] LIU Chengyan* and WANG Zhaomin, , 2014: On the Response of the Global Subduction Rate to Global Warming in Coupled Climate Models, ADVANCES IN ATMOSPHERIC SCIENCES, 31, 211-218.  doi: 10.1007/s00376-013-2323-9
    [2] FANG Changfang*, WU Lixin, and ZHANG Xiang, 2014: The Impact of Global Warming on the Pacific Decadal Oscillation and the Possible Mechanism, ADVANCES IN ATMOSPHERIC SCIENCES, 31, 118-130.  doi: 10.1007/s00376-013-2260-7
    [3] Kexin LI, Fei ZHENG, Jiang ZHU, Qing-Cun ZENG, 2024: El Niño and the AMO Sparked the Astonishingly Large Margin of Warming in the Global Mean Surface Temperature in 2023, ADVANCES IN ATMOSPHERIC SCIENCES.  doi: 10.1007/s00376-023-3371-4
    [4] ZHAO Qiaohua, SUN Jihua, ZHU Guangwei, 2012: Simulation and Exploration of the Mechanisms Underlying the Spatiotemporal Distribution of Surface Mixed Layer Depth in a Large Shallow Lake, ADVANCES IN ATMOSPHERIC SCIENCES, 29, 1360-1373.  doi: 10.1007/s00376-012-1262-1
    [5] Xinyi XING, Xianghui FANG, Da PANG, Chaopeng JI, 2024: Seasonal Variation of the Sea Surface Temperature Growth Rate of ENSO, ADVANCES IN ATMOSPHERIC SCIENCES, 41, 465-477.  doi: 10.1007/s00376-023-3005-x
    [6] MA Xiaoyan, GUO Yufu, SHI Guangyu, YU Yongqiang, 2004: Numerical Simulation of Global Temperature Change during the 20th Century with the IAP/LASG GOALS Model, ADVANCES IN ATMOSPHERIC SCIENCES, 21, 227-235.  doi: 10.1007/BF02915709
    [7] QIAN Cheng, FU Congbin, Zhaohua WU, YAN Zhongwei, 2011: The Role of Changes in the Annual Cycle in Earlier Onset of Climatic Spring in Northern China, ADVANCES IN ATMOSPHERIC SCIENCES, 28, 284-296.  doi: 10.1007/s00376-010-9221-1
    [8] Ran LIU, Changlin CHEN, Guihua WANG, 2016: Change of Tropical Cyclone Heat Potential in Response to Global Warming, ADVANCES IN ATMOSPHERIC SCIENCES, 33, 504-510.  doi: 10.1007/s00376-015-5112-9
    [9] QIAN Cheng, YAN Zhongwei, Zhaohua WU, FU Congbin, TU Kai, 2011: Trends in Temperature Extremes in Association with Weather-Intraseasonal Fluctuations in Eastern China, ADVANCES IN ATMOSPHERIC SCIENCES, 28, 297-309.  doi: 10.1007/s00376-010-9242-9
    [10] YUAN Zhuojian, QIAN Yu-Kun, QI Jindian, WU Junjie, 2012: The Potential Impacts of Warmer-Continent-Related Lower-Layer Equatorial Westerly Wind on Tropical Cyclone Initiation, ADVANCES IN ATMOSPHERIC SCIENCES, 29, 333-343.  doi: 10.1007/s00376-011-1100-x
    [11] Xiaofei GAO, Jiawen ZHU, Xiaodong ZENG, Minghua ZHANG, Yongjiu DAI, Duoying JI, He ZHANG, 2022: Changes in Global Vegetation Distribution and Carbon Fluxes in Response to Global Warming: Simulated Results from IAP-DGVM in CAS-ESM2, ADVANCES IN ATMOSPHERIC SCIENCES, 39, 1285-1298.  doi: 10.1007/s00376-021-1138-3
    [12] YANG Jing, BAO Qing, WANG Xiaocong, 2013: Intensified Eastward and Northward Propagation of Tropical Intraseasonal Oscillation over the Equatorial Indian Ocean in a Global Warming Scenario, ADVANCES IN ATMOSPHERIC SCIENCES, 30, 167-174.  doi: 10.1007/s00376-012-1260-3
    [13] WEI Xiaolin, LIU Qian, Ka Se LAM, WANG Tijian, 2012: Impact of Precursor Levels and Global Warming on Peak Ozone Concentration in the Pearl River Delta Region of China, ADVANCES IN ATMOSPHERIC SCIENCES, 29, 635-645.  doi: 10.1007/s00376-011-1167-4
    [14] Yiyong LUO, Jian LU, Fukai LIU, Xiuquan WAN, 2016: The Positive Indian Ocean Dipole-like Response in the Tropical Indian Ocean to Global Warming, ADVANCES IN ATMOSPHERIC SCIENCES, 33, 476-488.  doi: 10.1007/s00376-015-5027-5
    [15] Jingchao LONG, Suping ZHANG, Yang CHEN, Jingwu LIU, Geng HAN, 2016: Impact of the Pacific-Japan Teleconnection Pattern on July Sea Fog over the Northwestern Pacific: Interannual Variations and Global Warming Effect, ADVANCES IN ATMOSPHERIC SCIENCES, 33, 511-521.  doi: 10.1007/s00376-015-5097-4
    [16] Yawen DUAN, Peili WU, Xiaolong CHEN, Zhuguo MA, 2018: Assessing Global Warming Induced Changes in Summer Rainfall Variability over Eastern China Using the Latest Hadley Centre Climate Model HadGEM3-GC2, ADVANCES IN ATMOSPHERIC SCIENCES, 35, 1077-1093.  doi: 10.1007/s00376-018-7264-x
    [17] Tim LI, ZHANG Lei, Hiroyuki MURAKAMI, 2015: Strengthening of the Walker Circulation under Global Warming in an Aqua-Planet General Circulation Model Simulation, ADVANCES IN ATMOSPHERIC SCIENCES, 32, 1473-1480.  doi: 10.1007/s00376-015-5033-7
    [18] Chengyan LIU, Zhaomin WANG, Bingrui LI, Chen CHENG, Ruibin XIA, 2017: On the Response of Subduction in the South Pacific to an Intensification of Westerlies and Heat Flux in an Eddy Permitting Ocean Model, ADVANCES IN ATMOSPHERIC SCIENCES, 34, 521-531.  doi: 10.1007/s00376-016-6021-2
    [19] Yuan WANG, 2015: Air Pollution or Global Warming: Attribution of Extreme Precipitation Changes in Eastern China——Comments on "Trends of Extreme Precipitation in Eastern China and Their Possible Causes", ADVANCES IN ATMOSPHERIC SCIENCES, 32, 1444-1446.  doi: 10.1007/s00376-015-5109-4
    [20] SONG Yi, YU Yongqiang, LIN Pengfei, 2014: The Hiatus and Accelerated Warming Decades in CMIP5 Simulations, ADVANCES IN ATMOSPHERIC SCIENCES, 31, 1316-1330.  doi: 10.1007/s00376-014-3265-6

Get Citation+

Export:  

Share Article

Manuscript History

Manuscript received: 30 April 2015
Manuscript revised: 16 October 2015
Manuscript accepted: 27 October 2015
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Changes in Mixed Layer Depth and Spring Bloom in the Kuroshio Extension under Global Warming

  • 1. Physical Oceanography Laboratory/Qingdao Collaborative Innovation Center of Marine Science and Technology, Key Laboratory of Ocean-Atmosphere Interaction and Climate in Universities of Shandong, Ocean University of China, Qingdao 266100
  • 2. Scripps Institution of Oceanography, University of California San Diego, La Jolla, California, 92093, USA

Abstract: The mixed layer is deep in January-April in the Kuroshio Extension region. This paper investigates the response in this region of mixed layer depth (MLD) and the spring bloom initiation to global warming using the output of 15 models from CMIP5. The models indicate that in the late 21st century the mixed layer will shoal, and the MLD reduction will be most pronounced in spring at about 33°N on the southern edge of the present deep-MLD region. The advection of temperature change in the upper 100 m by the mean eastward flow explains the spatial pattern of MLD shoaling in the models. Associated with the shoaling mixed layer, the onset of spring bloom inception is projected to advance due to the strengthened stratification in the warming climate.

1. Introduction
  • The ocean mixed layer is a surface layer of vertically uniform temperature, salinity, and density, as a result of direct interaction with the atmosphere. The mixed layer depth (MLD) is determined by wind stirring, surface buoyancy forcing (i.e., freshwater and heat flux), and ocean circulation changes (Kraus and Businger, 1967). MLD is one of the most important quantities of the upper ocean, crucial to the substance exchange across the air-sea interface (e.g., heat flux, surface wave propagation, and ocean biological processes). Variability in oceanic uptake of atmospheric CO2 and SST is influenced by the mixed layer changes (Kraus and Businger, 1995). The mixed layer also controls the ocean's absorption of light and utilization of nutrients, two important factors affecting phytoplankton dynamics, and consequently, biological productivity in the ocean (Sverdrup, 1953; Yentsch, 1990).

    Much attention has been given to the seasonal cycle of MLD. (Kara et al., 2003) described the general features of the seasonal variation of MLD over the world. In the Kuroshio extension (KE) region, MLD is deep from January to April, shoals in summer, and deepens again in winter. (Qiu and Kelly, 1993) used a three-dimensional bulk mixed layer model of (Kraus and Turner, 1967) and studied the heat balance of the mixed layer over the KE. They showed that the variability of the heat flux drives the seasonal cycle of MLD. The model uses an assumption that the MLD can be estimated by the heat flux, freshwater flux, and wind stress at the air-sea interface, but neglects horizontal advection. Ocean dynamics plays a critical role in the KE system (Wu et al., 2003; Qiu et al., 2007; Taguchi et al., 2007; Xu et al., 2014). The western boundary currents carry warm water to the midlatitudes, releasing a great quantity of heat and moisture there to heat the atmosphere (Wu et al., 2012). In the KE there is warm temperature advection throughout the year. Since the warm advection makes the ocean lose heat to the atmosphere, it causes the surface density to increase and deepen the MLD. In the heat budget of the mixed layer, ocean heat Ekman transport and geostrophic advection plays a prominent role (Yim et al., 2013). As the ocean circulation changes under global warming, it most likely impacts the spatial distributions of SST and MLD, especially in the Kuroshio current and its extension (Sakamoto et al., 2005; Sato et al., 2006).

    MLD change also affects ocean biological process. Spring bloom refers to the rapid increase in phytoplankton abundance that commonly occurs in the early spring. During winter, wind-driven turbulence and surface cooling allow vertical mixing to replenish nutrients from depth to the mixed layer. Phytoplankton uses these nutrients for photosynthesis. Yet vertical mixing also causes high phytoplankton losses when phytoplankton's respiration exceeds primary production below the euphotic zone. For this reason, reduced illumination during winter limits phytoplankton's growth rates. In the spring, more light becomes available and stratification of the water column strengthens. As a result, vertical mixing is suppressed and phytoplankton and nutrients are kept near the surface, which promotes primary production. The definition and mechanism of spring bloom are discussed in (Mann and Lazier, 2005). The increasing primary production causes a strong growth of phytoplankton in spring. (Sverdrup, 1953) showed that there must exist a critical depth that blooming can only occur if the depth of mixed layer is less than the critical value. The critical depth was defined as a hypothetical surface mixing depth at which the integrated net growth rate over the water column becomes zero. Recently, studies (Behrenfeld, 2010; Boss and Behrenfeld, 2010; Taylor and Ferrari, 2011; Chiswell et al., 2015) have called this classic work into question, as the spring bloom is not solely caused by the shoaling mixed layer. (Chiswell, 2011) stated that Sverdrup's assumption of an evenly mixed phytoplankton layer was not applicable in most cases. (Huisman et al., 1999) used observational data to put forward a critical turbulence hypothesis that phytoplankton can bloom near the surface within a deep mixed layer if vertical mixing is low enough. (Chiswell, 2011) proposed an onset of stratification hypothesis that the spring bloom develops in weakly stratified layers. The initiation of spring bloom will be examined with observations in the context of these hypotheses.

    Under global warming, ocean and atmospheric circulations are projected to change significantly (Vecchi and Soden, 2007; Xie et al., 2010). MLD would also change because of circulation changes and increased thermal stratification. The winter MLD is generally projected to decrease (Luo et al., 2009). The MLD shoals as the anthropogenic warming is surface intensified, affecting mode waters in the North Pacific Ocean (Luo et al., 2009; Xu et al., 2012). (Xie et al., 2010) showed that the circulation change due to weakened mode water formation is more important than local atmospheric heat flux for SST variations in the subtropical gyre of the North Pacific.

    GCMs are an important tool to investigate the changes of MLD and the impact on biological processes under global warming. Most previous results were based on only one single coupled model and need to be verified in other models. Here, we take a multi-model approach to address the following questions: How does the MLD change under global warming? How do the CMIP5 results compare with previous studies? Besides a general shoaling under global warming, does the seasonal cycle of MLD also change? Does the inception of spring bloom start earlier? We will show that under global warming, MLD shoals and changes its seasonal cycle. This study considers both atmospheric and oceanic variables in the KE and examines the dominant mechanisms for the MLD changes. We diagnose the start time of the spring bloom in the KE and investigate how it changes in response to global warming. As the mixed layer shoals under global warming, the ability of the ocean to draw nutrient-rich water to the surface is reduced.

    The rest of the paper is arranged as follows: Section 2 describes the data and methods. Section 3 investigates the MLD changes under global warming, while section 4 diagnoses which variable is more important for the MLD change. Section 5 discusses the relationship between the mixed layer change and the spring bloom. Section 6 is a summary with discussion.

2. Data and methods Data
  • This paper uses the output from 15 CMIP5 coupled climate models (Table 1), which offer a multimodel perspective of simulated climate change and variability (Taylor et al., 2012). Both the historical (20th century with all forcing) simulation and the RCP4.5 scenario run (radiative forcing of 4.5 W m-2 by the year 2100, relative to preindustrial conditions) are used. The model output was obtained from the PCMDI at the Lawrence Livermore national laboratory.

    The resolution of atmospheric and oceanic variables is different within the same model and varies between models. We interpolated them on a 1°× 1° grid. Both the ensemble mean and the differences among models are investigated in the paper. We focus on the MLD variability and oceanic dynamics. The present-day climatology is based on the time average from 1951 to 2000 in the historical run, while the future mean state is calculated from 2051 to 2100 in the RCP4.5 run. A 50-year period is believed to be long enough to filter out the interannual variability. The change due to global warming is defined as the future mean state (RCP4.5 run, 2051-2100) minus the present-day climatology (historical run, 1951-2000). We only examine one member run of each model. The average of all models is defined as the ensemble mean. For example, we first calculate the MLD in each model, and then average for 15 models. The methods used here to process the data are the same as those employed by (Xu et al., 2012).

    For studying the spring bloom, we use the MODIS ocean color data (available from http://oceancolor.gsfc.nasa.gov/), including the daily chlorophyll-a and the photosynthetically active radiation, mapped at a resolution of 9 km, from 2004 to 2009. The daily net heat flux data from 2004 to 2009 is derived from the WHOI's OAFlux project (Yu et al., 2006). The weekly temperature and salinity data were downloaded from the China Argo real-time data center (http://www.argo.org.cn/). All the observational results presented in this paper are the mean state from 2004 to 2009. Only four models (CanESM2, GFDL-ESM2M, IPSL-CM5A-LR, MPI-ESM-LR) out of the 15 have both chlorophyll-a data and daily data of MLD and radiation. We use these four models in the study of the spring bloom.

  • There are various methods for determining MLD. Here, we regard both salinity and temperature as having effects on stratification. We use a fixed density criterion to calculate MLD. The MLD is defined as the depth where the increase in density from the surface value equals 0.03 kg m-3 (Xu et al., 2012). The vertical structure of the upper ocean in the KE was examined by 30 randomly selected profiles (not shown here), and it was found that this density criterion of 0.03 kg m-3 is suitable for defining the MLD.

    The net heat flux is the sum of longwave radiation, shortwave radiation, latent heat flux, and sensible heat flux at the surface. The temperature advection, u(∂ T/∂ x)+v(∂ T/∂ y), in the mixed layer is calculated as the temperature advection at a depth of 50 m. In this formula, u,v are the horizontal velocity components, T is the SST, ∂/∂ x+∂/∂ y is the horizontal gradient operators.

3. Seasonal change of MLD
  • This section studies changes in the seasonal cycle of MLD in response to global warming. The analysis region is (25°-45°N, 131°E-160°W). Figure 1a shows the present-day winter (January-March) climatological MLD (1951-2000). The MLD in the KE is deep and its maximum exceeds 200 m. According to (Xu et al., 2014), the deep mixed layer near the KE shows significant differences between observations and current climate models. In observations, there are two MLD maxima deeper than 150 m to the north and south of the KE, respectively. Sandwiched between is a shallower mixed layer along the KE jet. By contrast, CMIP5 models do not capture this feature, with one single broad pool of deep MLD and a sharp MLD front to the south that slants northeastward. The present study will only focus on the deep MLD changes as a whole, ignoring its detailed structures.

    Figure 1.  (a) Present-day winter (January-March) climatology (1951-2000) of MLD (colored scale bar; units: m). (b) Seasonal cycle of the MLD for present-day climatology (1951-2000) and its change [RCP4.5 run (2051-2100) minus historical run (1951-2000)]. The zonal mean (135$^\circ$-175$^\circ$E) MLD is shown by black contours in intervals of 50 m, and the MLD change is shown by the coloring in units of m. (c, d) Historical mean MLD (1951-2000; colored scale bar) and 21st century mean MLD (2051-2100; blue contours) in (c) March and (d) April [MLD change is superimposed (black dotted contours at 10 m intervals)].

    Figure 1b shows the seasonal cycle of the zonal mean (135°-175°E) MLD for the present-day climatology and its change under global warming. The seasonal cycle of MLD in present-day climatology is obvious: the mixed layer is deep in January-April, and the deepest MLD (about 250 m) is located at about 33°N; the MLD shoals after April, to less than 50 m, and then deepens again in November-December.

    The MLD change under global warming shows seasonal variability. The biggest change, almost -40 m, takes place in April. The maximum change of MLD under global warming is collocated with the maximum MLD in the present-day climatology in April at 33°N. The MLD change in March is less than that in April. The pattern of the MLD change is different between March and April: the MLD change in March peaks near the northeast-slanted MLD front of the present-day climatology, while the April change is flatter (Figs. 1c and d). While the ensemble mean results show a clear seasonality in MLD change in response to global warming, different models may have different characteristics. Figure 2 shows the inter-model standard deviation of winter mean MLD among the 15 models. Large inter-model bias (>25 m) appears between 25°N and 30°N, to the south of the winter deep mixed layer front. But the difference is relatively small (<15 m) in the deep mixed layer region, where it shows a significant seasonal change of MLD (>30 m).

    All of the 15 models show a shoaling MLD in response to global warming. Ten models (BCC_CSM1.1, CanESM2, CCSM4, HadGEM2-CC, MIROC5, MIROC-ESM, MPI-ESM-LR, CMCC-CM, CSIRO Mk3.6.0, NorESM1-M) have similar results as the ensemble mean, but the other five (ACCESS1.3, GFDL-CM3, GFDL-ESM2M, IPSL-CM5A-LR, MRI-CGCM3) do not show similar MLD changes. In these five models, the maximum change of MLD under global warming is collocated with the maximum MLD in the present-day climatology in both March and April. The MLD change in March is more than that in April. The reason why these five models do not have the same characteristics needs further research but will not be discussed in this paper. We focus on the ensemble mean results in the following sections.

    Figure 2.  Inter-model standard deviation (colored scale bar; units: m) of the winter mean MLD for January-March.

4. Atmospheric and oceanic effects on MLD change
  • This section investigates why MLD changes under global warming. Following (Qiu and Kelly, 1993), we diagnose atmospheric and oceanic factors influencing MLD, including the important role of the western boundary currents in the KE.

  • Figure 3 shows the changes of MLD and heat flux in March and April in response to global warming, with positive meaning the ocean is absorbing more heat. The heat flux change varies from -15 W m-2 to 15 W m-2. In March, the ocean tends to release more heat in most of the area, but it absorbs more heat in a long and narrow strip between 25°N and 30°N (Fig. 3a). In April, the ocean absorbs more heat from the atmosphere, and the area of negative heat flux becomes smaller (Fig. 3b). The ocean absorbing more heat from the atmosphere in April than March explains why the MLD shoals more in April under global warming. However, the heat flux change under global warming in March and April cannot explain the spatial pattern of the MLD change.

  • In the KE, ocean currents are strong and play an important role in the formation of the deep mixed layer (Wu et al., 2012; Yim et al., 2013). The western boundary currents carry warm water to the midlatitudes. The warm advection makes the ocean lose a large quantity of heat to the atmosphere, causing a deep MLD. Figure 4 shows the change of temperature advection in the mixed layer in March and April in response to global warming. The shoaling MLD seems spatially well correlated with the warm advection in the mixed layer both in March and April. A warm mixed layer temperature advection shoals the MLD because it enhances stratification.

    Figure 3.  Change in net heat flux (future flux minus present-day flux; colored scale bar; units: W m$^-2$), where positive means increased heat into the ocean, for (a) March and (b) April. MLD change is superimposed (black dotted contours at 10 m intervals).

    Figure 4.  Future minus present-day change in temperature advection (colored scale bar; units: 10$^-8$ $^\circ$C s$^-1$) in (a) March and (b) April. Negative values mean warm advection. MLD change is superimposed (black dotted contours in 10 m intervals).

    Figure 5.  Temperature advection in March (colored scale bar; units: 10$^-8$ $^\circ$C s$^-1$) (a) by holding potential temperature constant in the present day and changing ocean currents under global warming, and (b) by holding ocean currents constant in the present day and changing potential temperature under global warming. MLD change is superimposed (black contours at 10 m intervals).

    We further investigate which term dominates the temperature advection: changes in potential temperature, or ocean currents? By dividing velocity and potential temperature into climatology and anomalies, \(u=\overline u+u'\), \(v=\overline v+v'\), and \(T=\overline T+T'\), we decompose the temperature advection into the following components: \begin{eqnarray} &&u\dfrac{\partial T}{\partial x}+v\dfrac{\partial T}{\partial y}\nonumber\\ &=&(\overline {u}+u')\dfrac{\partial\overline {T}+T'}{\partial x}+(\overline {v}+v')\dfrac{\partial\overline {T}+T'}{\partial y}\nonumber\\ &=&\overline {u}\dfrac{\partial\overline {T}}{\partial x}+\overline {v}\dfrac{\partial\overline {T}}{\partial y}+ \left(u'\dfrac{\partial\overline {T}}{\partial x}+v'\dfrac{\partial\overline {T}}{\partial y}\right)\!+\! \left(\overline {u}\dfrac{\partial T'}{\partial x}\!+\!\overline {v}\dfrac{\partial T'}{\partial y}\right),\quad\quad \end{eqnarray} where \(\overline T,\overline u\) and \(\overline v\) are potential temperature and the zonal and meridional currents in the present-day climatology, respectively, and T',u' and v' are their changes under global warming. One keeps potential temperature constant at the present-day climatology but changes currents, while the other holds currents constant but allows potential temperature to vary as climate warms. Figure 5 shows the result in March. If we hold the potential temperature unchanged, large changes in temperature advection take place off the Japanese coast. On the other hand, the advection of anomalous temperature by the mean current produces the spatial pattern of total advection. In March and April, both potential temperature and ocean currents are important for temperature advection and have impacts on the change of MLD under global warming. However, in the region of maximum MLD change on the southern flank of the deep mean-MLD region, the potential temperature change plays a dominant role in MLD change.

    Figure 6 shows the longitude-depth section of temperature advection change and potential temperature change under global warming, averaged in 25°-30°N in March. The maximum warm advection is in the 100 m upper layer, with a maximum from 150°E to 180°E. The maximum increase of potential temperature is in the 100 m upper layer, with a maximum near 150°E. East of 150°E, temperature warming decreases gradually, but the eastward background currents cause a warm advection. At 170°E, where both the anomalous temperature gradient and mean currents are strong, the mixed layer shoals the most. The distribution of potential temperature change below 100 m is small.

5. Spring blooms under global warming
  • The shoaling mixed layer under global warming has an important influence on the ecosystem. Many variables, such as organic carbon, nitrate, nitrite, and chlorophyll, can affect phytoplankton production (Yentsch, 1990). This section investigates the spring blooms under global warming, based on the CMIP5 models. Having examined various observational indicators of spring bloom, we focus here on surface chlorophyll, for which satellite data are readily available.

    From autumn to winter, the ocean loses heat to the atmosphere, and the mixed layer gradually deepens. Strong mixing happens throughout the whole depth of the mixed layer, entraining new nutrients into the mixed layer and leading to an increase in production. In winter, the MLD reaches its deepest value at t1 and the convective overturn is strong (Fig. 7a). At the end of winter, the slow convective overturn cannot maintain a deep mixed layer, with reduced turbulence; the mixed layer begins to shoal slightly, lightly increasing phytoplankton concentrations at the surface. The transition from strong mixing to low turbulence occurs at about t1 (Fig. 7a), the time (t1) after the mixed layer reaches its deepest point. After t1, the net heat flux begins to rise (Fig. 7b), which means that the ocean loses less heat to atmosphere, and after t2 it starts to absorb heat from the atmosphere. Meanwhile, the MLD starts to shoal sharply and the rate of increase in surface chlorophyll begins to accelerate. Figure 7c shows the buoyancy frequency, defined as N=[-g(∂ρ/∂ z)/ρ]1/2, where g is the gravitational acceleration, z denotes geometric height and ρ is the potential density. The sharp increase in N acts as a barrier for the downward turbulence generated in the surface mixed layer by atmospheric disturbances (Gill, 1982). At the time of t1, the MLD starts to shoal and surface chlorophyll begins to increase, but the net heat flux at the surface remains negative (upward). A weak stratification supports a weak spring bloom at the surface. At the time of t2, surface chlorophyll increases to the second peak and the net heat flux turns positive, while MLD has already shoaled sharply with N continuing to increase. A strong stratification supports an intense bloom. The net heat flux is not the only determinant of MLD in the KE (Qiu et al., 2007; Taguchi et al., 2007), so there is a 60-day lag between t1 and t2. Thus, overall, the MLD shoaling is a good indicator for the timing of the spring bloom, and the time of zero net heat flux is a sign to expedite the spring bloom in the KE.

    Figure 6.  Longitude-depth section of the meridional mean (25$^\circ$-30$^\circ$N) change in temperature advection (colored scale bar; units: 10$^-8$ $^\circ$C s$^-1$) in March. Temperature change (black lines; units: $^\circ$C) and currents (black arrows; m s$^-1$) of present-day climatology (1951-2000) are superimposed.

    Figure 7.  Regional average (30$^\circ$-40$^\circ$N, 140$^\circ$-160$^\circ$E) of (a) MLD (blue line; units: m) and surface chlorophyll concentration (green line; units: mg m$^-3$) from satellite data, (b) net heat flux from OAFlux (blue line; units: W m$^-2$; negative means ocean is losing heat) and chlorophyll concentration (green line; units: mg m$^-3$), and (c) buoyancy frequency (s$^-1$) from Argo. The dashed vertical lines mark the first and second chlorophyll blooms. The time axis starts on 1 January.

    Figure 8 compares the climatological distribution of surface chlorophyll concentration between observations and the models. The model simulations capture the overall spatial pattern but are biased (too high) in terms of surface chlorophyll concentration, especially in the midlatitudes (35°-45°N).

    Figure 9 compares the daily variation of surface chlorophyll concentration at present and under global warming. Compared to the current climatology, the spring increase in chlorophyll shows a tendency to start earlier by about 10 days under global warming in the KE, accompanied by a decrease in MLD due to the surface warming trend (Fig 9a). In Fig. 9b, surface chlorophyll increases to the second peak earlier by about 15 days under global warming, with the net heat flux turning positive ahead of time (Fig. 9b). In the CMIP5 models, the shoaling trend of MLD is consistent with the early onset of the spring bloom under global warming.

    Figure 8.  (a) Satellite-derived chlorophyll distribution (units: mg m$^-3$) obtained from MODIS for 2004-09. (b) Simulated climatological distribution of chlorophyll (units: mg m$^-3$) in the historical simulation. (c) Differences in chlorophyll concentration (units: mg m$^-3$; simulated minus observation).

    Figure 9.  Daily evolution of the regional averaged (30$^\circ$-40$^\circ$N, 140$^\circ$-160$^\circ$E) (a) MLD (black lines; units: m) and chlorophyll concentration (green lines; mg m$^-3$) at the surface (dashed vertical lines mark the first chlorophyll blooms in the historical and RCP4.5 simulations), and (b) net heat flux (black line; units: W m$^-2$; negative means the ocean is losing heat) and chlorophyll concentration (green lines; units: mg m$^-3$) at the surface (dashed vertical lines mark the second chlorophyll blooms in the historical and RCP4.5 simulations). Note that the results are the ensemble mean.

6. Summary and discussion
  • This paper examines the change in the seasonal cycle of MLD in the KE in response to global warming, based on the output of 15 CMIP5 models. The MLD becomes shallower, especially in March and April. Under global warming, the MLD in April shoals mostly in the region where the MLD is presently large; while in March, it shoals the most in the region of the steep mean MLD gradient. This characteristic of MLD change varies somewhat among models and further studies are needed to determine the factors responsible for the different model behavior.

    Changes in both surface heat flux and ocean warm temperature advection contribute to the spring shoaling of the mixed layer. The advection of temperature change in the upper 100 m by the mean eastward current explains the spatial pattern of MLD change in spring. The spatial distribution of mixed layer temperature change——large in the west and reduced in the east——is the main reason for temperature advection change under global warming. This result highlights the importance of the ocean surface warming pattern (Xie et al., 2010). It also raises another question as to what causes this spatial distribution of potential temperature change. In the extra-tropics, wind stress forcing and ocean heat transport may be important. Diagnostic methods based on the mixed layer heat budget need to be developed to investigate the underlying mechanism of the temperature pattern formation.

    The triggering mechanisms for spring bloom in the KE were examined using satellite data, and it was found that the strengthened stratification and mixed layer shoaling can cause a surface bloom of chlorophyll, and the net heat flux turning to positive from negative also causes a second peak of chlorophyll in the current climate. Under global warming, model projections suggest an early onset of the spring phyto- plankton bloom——a change that is consistent with the shoaling of the mixed layer in the warming climate. This result still needs to be tested with observations and model output of the vertical dimension of chlorophyll concentration lacking of liability data. (Hashioka et al., 2009) obtained a similar result that the spring bloom initiates earlier by about 10 to 20 days under global warming, although the physical processes were not investigated in detail. The chlorophyll concentration can alter solar penetration through the ocean (Nakamoto et al., 2001; Murtugudde et al., 2002). Thus, there may be feedback between physical and biological changes.

    The total radiative forcing begins to stabilize around 2070 under RCP4.5 (Taylor et al., 2012). The ocean response comprises a fast response of the mixed layer warming (10-year timescale, approximately) and a slow response involving the deeper ocean (Held et al., 2010). The fast response dominates as the radiative forcing increases, while the slow response takes over after the radiative forcing has stabilized. The fast response is associated with increased upper-ocean stratification and shoaling of the mixed layer, as discussed here. The slow response is associated with a slightly reduced upper ocean stratification (Long et al., 2014) and a weak increase in MLD in the KE (Xu et al., 2013), despite the continued increase in surface temperature. In future work, we intend to investigate the distinct fast and slow responses in the spring bloom.

Reference

Catalog

    /

    DownLoad:  Full-Size Img  PowerPoint
    Return
    Return