Advanced Search
Article Contents

Response of North Pacific Eastern Subtropical Mode Water to Greenhouse Gas Versus Aerosol Forcing


doi: 10.1007/s00376-015-5092-9

  • Mode water is a distinct water mass characterized by a near vertical homogeneous layer or low potential vorticity, and is considered essential for understanding ocean climate variability. Based on the output of GFDL CM3, this study investigates the response of eastern subtropical mode water (ESTMW) in the North Pacific to two different single forcings: greenhouse gases (GHGs) and aerosol. Under GHG forcing, ESTMW is produced on lighter isopycnal surfaces and is decreased in volume. Under aerosol forcing, in sharp contrast, it is produced on denser isopycnal surfaces and is increased in volume. The main reason for the opposite response is because surface ocean-to-atmosphere latent heat flux change over the ESTMW formation region shoals the mixed layer and thus weakens the lateral induction under GHG forcing, but deepens the mixed layer and thus strengthens the lateral induction under aerosol forcing. In addition, local wind changes are also favorable to the opposite response of ESTMW production to GHG versus aerosol.
  • 加载中
  • Donner, L. J., Coauthors , 2011: The dynamical core,physical parameterizations, and basic simulation characteristics of the atmospheric component AM3 of the GFDL global coupled model CM3. J. Climate, 24, 3484-3519, doi: 10.1175/2011 JCLI3955.1.625abeb6-c951-4f2b-9979-2a30ddd1605e0c3f00612575944ce10dfd22aa8048bbhttp%3A%2F%2Fwww.osti.gov%2Fscitech%2Fservlets%2Fpurl%2F1117959refpaperuri:(3cfd733b206c59475a8af1c78a7c55af)http://www.osti.gov/scitech/servlets/purl/1117959
    Griffies, S. M., Coauthors , 2011: The GFDL CM3 coupled climate model: Characteristics of the ocean and sea ice simulations. J. Climate, 24, 3520-3544, doi: 10.1175/2011JCLI 3964.1.10.1175/2011JCLI3964.1cf90970350b57f84cd422b0676b9f341http%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F2011JCli...24.3520Ghttp://adsabs.harvard.edu/abs/2011JCli...24.3520GABSTRACT This paper documents time mean simulation characteristics from the ocean and sea ice components in a new coupled climate model developed at the NOAA Geophysical Fluid Dynamics Laboratory (GFDL). The GFDL Climate Model version 3 (CM3) is formulated with effectively the same ocean and sea ice components as the earlier CM2.1 yet with extensive developments made to the atmosphere and land model components. Both CM2.1 and CM3 show stable mean climate indices, such as large-scale circulation and sea surface temperatures (SSTs). There are notable improvements in the CM3 climate simulation relative to CM2.1, including a modified SST bias pattern and reduced biases in the Arctic sea ice cover. The authors anticipate SST differences between CM2.1 and CM3 in lower latitudes through analysis of the atmospheric fluxes at the ocean surface in corresponding Atmospheric Model Intercomparison Project (AMIP) simulations. In contrast, SST changes in the high latitudes are dominated by ocean and sea ice effects absent in AMIP simulations The ocean interior simulation in CM3 is generally warmer than in CM2.1, which adversely impacts the interior biases.
    Han W. Q., G. A. Meehl, and A. X. Hu, 2006: Interpretation of tropical thermocline cooling in the Indian and Pacific oceans during recent decades. Geophys. Res. Lett., 33,L23615, doi: 10.1029/2006GL027982.
    Hanawa K., L. D. Talley, 2001: Mode waters. Ocean Circulation and Climate, G. Siedler, J. Church, and J. Gould, Eds., International Geophysical Series, Vol. 77, Academic Press, 373- 400.
    Hautala S. L., D. H. Roemmich, 1998: Subtropical mode water in the northeast Pacific Basin. Journal of Geophysical Research: Oceans, 103, 13 055- 13 066.10.1029/98JC0101594b11d1757d54d6995e002109d8cb9d4http%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1029%2F98JC01015%2Fpdfhttp://onlinelibrary.wiley.com/doi/10.1029/98JC01015/pdfA new type of mode water in the upper thermocline of the eastern subtropical North Pacific is identified and examined using data from World Ocean Circulation Experiment high-resolution repeat expendable bathythermograph (XBT) section PX37 and archives of historical XBT data. This water mass (labeled Eastern Subtropical Mode Water) is characterized by a subsurface potential vorticity minimum and is located east of Hawaii (Northeast Pacific Basin) in a density range of 24–25.4 σ θ . It is a distinct water mass from the classical subtropical mode water (STMW) of the western Pacific. Eastern STMW is formed as a relatively deep late-winter mixed layer, associated with the subtropical/subpolar water mass boundary near 25°–30°N, 135°–140°W, and is capped and subducted into the permanent thermocline. Along a section between San Francisco and Honolulu, Eastern STMW production is seen in every year for which there is adequate data. In this section the volume of Eastern STMW formed each winter and the temperature of the potential vorticity minimum are similar during the periods 1970–1979 and 1991–1997.
    Hosoda S., S. P. Xie, K. Takeuchi, and M. Nonaka, 2001: Eastern North Pacific subtropical mode water in a general circulation model: Formation mechanism and salinity effects. J. Geophys. Res., 106, 19 671- 19 681.10.1029/2000JC0004434bc96bef2ed448a44f6423dcb06fd92ahttp%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1029%2F2000JC000443%2Fabstracthttp://onlinelibrary.wiley.com/doi/10.1029/2000JC000443/abstractThe Eastern North Pacific Subtropical Mode Water (ESTMW) is a water mass of low potential vorticity (PV) and appears as a weak pycnostad or thermostad. Distinct from other subtropical mode waters, it forms in the absence of a deep winter mixed layer. The formation mechanism of this ESTMW is investigated using an ocean general circulation model that is forced by monthly climatological temperature, salinity, and wind stress at the sea surface. An equation based on the ventilated thermocline theory is used to diagnose the initial PV of a water mass right after its subduction. In this equation, three factors affect the initial PV: the spacing of density outcrop lines, the mixed layer depth gradient, and the vertical velocity at the bottom of mixed layer. Among them the wide spacing between outcrop lines is the most important for ESTMW's low PV instead of the deep mixed layer, which is most important for classical mode waters. It is found that weak gradients in both sea surface temperature and salinity in the direction of mixed layer flow are important for the low PV formation. A low-salinity tongue that extends southeastward off North America is responsible for the small surface density gradient in the eastern North Pacific and contributes to the formation of the ESTMW. An additional experiment forced with observed freshwater flux demonstrates that the southward advection of fresher water from the high latitude along the eastern boundary is the cause of this low-salinity tongue.
    Hu H. B., Q. Y. Liu, Y. Zhang, and W. Liu, 2011: Variability of subduction rates of the subtropical North Pacific mode waters. Chinese Journal of Oceanology and Limnology, 29, 1131- 1141.10.1007/s00343-011-0237-x1c65ed382f34507cc434281ad85a8ed3http%3A%2F%2Flink.springer.com%2F10.1007%2Fs00343-011-0237-xhttp://d.wanfangdata.com.cn/Periodical_zghyhzxb201105027.aspxThe climatology subduction rate for the entire Pacific is known,but the mechanism of interannual to decadal variation remains unclear.In this study,we calculated the annual subduction rates of three types of North Pacific subtropical mode waters using a general circulation model (LICOM 1.0)for the period of 1958-2001.The model experiments focused on interarnual variations of ocean dynamical processes under daily wind forcings and seasonal heat fluxes.The mode water formation region was defined by a potential vorticity minimum at outcrop locations.The model results show that two subduction rate maxima (锛100 m/a) were located in the Subtropical Mode Water (STMW) and the Central Mode Water (CMW) formation regions.These regions are consistent with a climatologically calculated value.The subduction rate in the Eastern Subtropical Mode Water (ESTMW) formation region was smaller at about 75 m/a.The subduction rate shows clear interannual and decadal variations associated with oceanic dynamic variabilities.The average subduction rate of the STMW was much smaller during the period of 1981-1990 compared with other periods,while that of the CMW had a negative anomaly before 1975 and a positive anomaly after 1978.The variability agreed with Ekman and geostrophic advections and mixed layer depths.The interannual variability of the subduction rate for the ESTMW was smallest during 1970-1990,as a result of a weak wind stress curl.This paper explores how interannual signals from the atmosphere are stored in different parts of the ocean,and thus may contribute to a better understanding of feedback mechanisms for the Pacific Decadal Oscillation (PDO) event.
    Huang R. X., B. Qiu, 1994: Three-dimensional structure of the wind-driven circulation in the subtropical North Pacific. J. Phys. Oceanogr., 24, 1608- 1622.10.1175/1520-0485(1994)024<1608:TDSOTW>2.0.CO;2f8937805d5f3473e6e2c36cad08d463dhttp%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F1994JPO....24.1608Hhttp://adsabs.harvard.edu/abs/1994JPO....24.1608HAbstract The subduction rate is calculated for the North Pacific based on Levitus climatology data and Hellerman and Rosenstein wind stress data. Because the period of effective subduction is rather short, subduction rates calculated in Eulerian and Lagrangian coordinates are very close. The subduction rate defined in the Lagrangian sense consists of two parts. The first part is due to the vertical pumping along the one-year trajectory, and the second part is due to the difference in the winter mixed layer depth over the one-year trajectory. Since the mixed layer is relatively shallow in the North Pacific, the vertical pumping term is very close to the Ekman pumping, while the sloping mixed layer base enhances subduction, especially near the Kuroshio Extension. For most of the subtropical North Pacific, the subduction rate is no more than 75 m yr 611 , slightly larger than the Ekman pumping. The water mass volume and total amount of ventilation integrated for each interval of 0.2σ unit is computed. The corresponding renewal time for each water mass is obtained. The inferred renewal time is 5–6 years for the shallow water masses (σ = 23.0–25.0), and about 10 years for the subtropical mode water (σ = 25.2–25.4). Within the subtropical gyre the total amount of Ekman pumping is 28.8 Sv (Sv ≡ 10 6 m 3 s 611 ) and the total subduction rate is 33.1 Sv, which is slightly larger than the Ekman pumping rate. To this 33.1 Sv, the vertical pumping contributes 24.1 Sv and the lateral induction 9 Sv. The maximum barotropic mass flux of the subtropical gyre is about 46 Sv (cut of 135°E). This mass flux is partitioned as follows. The total horizontal mass flux in the ventilated thermocline, the seasonal thermocline, and the Ekman layer is about 30 Sv, and the remaining 16 Sv is in the unventilated thermocline. Thus, about one-third of the man flux in the wind-driven gyre is sheltered from direct air–sea interaction.
    Inui T., K. Takeuchi, and K. Hanawa, 1999: A numerical investigation of the subduction process in response to an abrupt intensification of the westerlies. J. Phys. Oceanogr., 29, 1993- 2015.10.1175/1520-0485(1999)0292.0.CO;29dd0ad2b41137229fb709fd6f746f482http%3A%2F%2Fci.nii.ac.jp%2Fnaid%2F80011344500http://ci.nii.ac.jp/naid/80011344500Abstract A three-dimensional ocean general circulation model, forced by idealized zonal winds, is used to investigate the effect of an abrupt intensification of westerly winds on the subduction process. Four experiments are carried out: 1) a control experiment with standard wind stress forcing, 2) an intensified winds experiment with wind stress that is larger in the region of the westerlies than the control, 3) an increased surface cooling experiment, and 4) an experiment with both intensified wind stress and surface cooling. Experiments 2 through 4, which contain surface anomalous forcing, are run from the steady state obtained in experiment 1, the control experiment. The results obtained for each of these runs are compared to the control experiment. A subarctic tracer injection experiment is also carried out to verify the differences in the subduction process of each of these experiments. In the wind stress intensified experiment, an examination of the subsurface temperature field shows that negative temperature anomalies occupy the western portion of the southern part of the subtropical gyre, whereas in the surface cooling experiment, negative temperature anomalies occupy the eastern portion of the basin. The source of these negative temperature anomalies is not local since the forcing in the southern part of the subtropical gyre does not change from the control. A close analysis of the evolution of a subarctic surface tracer field indicates that the intensification of the wind stress increases the tracer concentrations, whereas surface cooling decreases the temperature in the region that contains the maximum tracer concentration. In the standard case, the mixed layer is deep (shallow) in the northern (southern) part of the subtropical gyre. Between these two regions a mixed layer front, where the mixed layer depth changes drastically from north to south, exists. A water column with low potential vorticity that originates in the mixed layer penetrates into a subsurface layer from the point where an outcrop line and the mixed layer front intersect. This point is called the penetration point. Intensified westerly winds bring about a deeper thermocline and shoaling subsurface isopycnals. These shoaling subsurface isopycnals are not predicted in classical models such as that of Luyten et al. The model experiment with intensified westerlies demonstrates that the penetration point shifts to the west. As a result, low potential vorticity water penetrates southwestward from the shifted penetration point and takes a more westward path. Therefore, the negative temperature anomalies appear in the southwestern part of the subtropical gyre. This study shows that the westward shift of the path of low potential vorticity water could cause the shoaling of subsurface isopycnal surfaces. The intensification of the westerlies increases Ekman pumping and cools the ocean surface by enhancing sensible and latent heat flux. In the surface cooling experiment, the position of the outcrop lines moves southward significantly. This southward shift makes the subducted water colder and distributes it throughout the ventilated region of the southern part of the subtropical gyre. The combined effect of wind intensification and surface cooling is approximately a linear combination of both experiments.
    Latif M., T. P. Barnett, 1994: Causes of decadal climate variability over the North Pacific and North America. Science, 266, 634- 637.10.1126/science.266.5185.634177934570dda648e0d1e61710b3fb3e761f0403dhttp%3A%2F%2Fonlinelibrary.wiley.com%2Fresolve%2Freference%2FPMED%3Fid%3D17793457http://med.wanfangdata.com.cn/Paper/Detail/PeriodicalPaper_PM17793457ABSTRACT The cause of decadal climate variability over the North Pacific Ocean and North America is investigated by the analysis of data from a multidecadal integration with a state-of-the-art coupled ocean-atmosphere model and observations. About one-third of the low-frequency climate variability in the region of interest can be attributed to a cycle involving unstable air-sea interactions between the subtropical gyre circulation in the North Pacific and the Aleutian low-pressure system. The existence of this cycle provides a basis for long-range climate forecasting over the western United States at decadal time scales.
    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.
    Marshall J. C., R. G. Williams, and A. J. G. Nurser, 1993: Inferring the subduction rate and period over the North Atlantic. J. Phys. Oceanogr., 23, 1315- 1329.10.1175/1520-0485(1993)023<1315:ITSRAP>2.0.CO;283938c57b0c78da7adbbcb644a19ce4ahttp%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F1993JPO....23.1315Mhttp://adsabs.harvard.edu/abs/1993JPO....23.1315MAbstract The annual rate at which mixed-layer fluid is transferred into the permanent thermocline hat is, the annual subduction rate S ann and the effective subduction period effs inferred from climatological data in the North Atlantic. From its kinematic definition, S ann is obtained by summing the vertical velocity at the base of the winter mixed layer with the lateral induction of fluid through the sloping base of the winter mixed layer. Geostrophic velocity fields, computed from the Levitus climatology assuming a level of no motion at 2.5 km, are used; the vertical velocity at the base of the mixed layer is deduced from observed surface Ekman pumping velocities and linear vorticity balance. A plausible pattern of S ann is obtained with subduction rates over the subtropical gyre approaching 100 m/yr wice the maximum rate of Ekman pumping. The subduction period eff is found by viewing subduction as a transformation process converting mixed-layer fluid into stratified thermocline fluid. The effective period is that period of time during the shallowing of the mixed layer in which sufficient buoyancy is delivered to permit irreversible transfer of fluid into the main thermocline at the rate S ann . Typically eff is found to be 1 to 2 months over the major part of the subtropical gyre, rising to 4 months in the tropics. Finally, the heat budget of a column of fluid, extending from the surface down to the base of the seasonal thermocline is discussed, following it over an annual cycle. We are able to relate the buoyancy delivered to the mixed layer during the subduction period to the annual-mean buoyancy forcing through the sea surface plus the warming due to the convergence of Ekman heat fluxes. The relative importance of surface fluxes (heat and freshwater) and Ekman fluxes in supplying buoyancy to support subduction is examined using the climatologist observations of Isemer and Hasse, Schmitt et al., and Levitus. The pumping down of fluid from the warm summer Ekman layer into the thermocline makes a crucial contribution and, over the subtropical gyre, is the dominant term in the thermodynamics of subduction.
    Mitchell J. F. B., T. C. Johns, J. M. Gregory, and S. F. B. Tett, 1995: Climate response to increasing levels of greenhouse gases and sulphate aerosols. Nature, 376, 501- 504.10.1038/376501a06a8f5df26618ac9b9fee8276ac6c031chttp%3A%2F%2Fwww.nature.com%2Fdoifinder%2F10.1038%2F376501a0http://www.nature.com/doifinder/10.1038/376501a0Presents a study which simulated past and future climate variations and responses to radiation-scattering by greenhouse gases and sulphate aerosols. Use of a coupled ocean-atmosphere general circulation model; Agreement between observed global mean and large-scale patterns of temperature; Prediction of future global mean warming of 0.3 Kelvin per decade for greenhouse gases alone.
    Nakamura H., 1996: A pycnostad on the bottom of the ventilated portion in the central subtropical North Pacific: Its distribution and formation. Journal of Oceanography, 52, 171- 188.e83089269ade0472cf521047594199cdhttp%3A%2F%2Flink.springer.com%2F10.1007%2FBF02235668/s?wd=paperuri%3A%28b4b9204a586becfe28c8375f47b94eee%29&filter=sc_long_sign&tn=SE_xueshusource_2kduw22v&sc_vurl=http%3A%2F%2Flink.springer.com%2F10.1007%2FBF02235668&ie=utf-8
    Rotstayn L. D., S. J. Jeffrey, M. A. Collier, S. M. Dravitzki, A. C. Hirst, J. I. Syktus, and K. K. Wong, 2012: Aerosol-induced changes in summer rainfall and circulation in the Australasian region: A study using single-forcing climate simulations. Atmospheric Chemistry & Physics, 12, 5107- 6377.ed896764a0de2a70c5a7945c7efc478chttp%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F2012ACPD...12.5107R/s?wd=paperuri%3A%28bf18d8e91e85e610b75000167f0c43e8%29&filter=sc_long_sign&tn=SE_xueshusource_2kduw22v&sc_vurl=http%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F2012ACPD...12.5107R&ie=utf-8
    Sen Gupta, S. A., N. C. Jourdain, J. N. Brown, D. Monselesan, 2013: Climate drift in the CMIP5 models. J.Climate, 26, 8597- 8615.10.1175/JCLI-D-12-00521.15026bc25-dcd4-4b81-be04-b0843c2044a4e1f1737f9ab760b9458a71a43e84b54bhttp%3A%2F%2Fadsabs.harvard.edu%2Fabs%2F2013JCli...26.8597Grefpaperuri:(b0439faedf2731b4221c87c98d4b9b9e)http://adsabs.harvard.edu/abs/2013JCli...26.8597GClimate models often exhibit spurious long-term changes independent of either internal variability or changes to external forcing. Such changes, referred to as model 'drift,' may distort the estimate of forced change in transient climate simulations. The importance of drift is examined in comparison to historical trends over recent decades in the Coupled Model Intercomparison Project (CMIP). Comparison based on a selection of metrics suggests a significant overall reduction in the magnitude of drift from phase 3 of CMIP (CMIP3) to phase 5 of CMIP (CMIP5). The direction of both ocean and atmospheric drift is systematically biased in some models introducing statistically significant drift in globally averaged metrics. Nevertheless, for most models globally averaged drift remains weak compared to the associated forced trends and is often smaller than the difference between trends derived from different ensemble members or the error introduced by the aliasing of natural variability. An exception to this is metrics that include the deep ocean (e.g., steric sea level) where drift can dominate in forced simulations. In such circumstances drift must be corrected for using information from concurrent control experiments. Many CMIP5 models now include ocean biogeochemistry. Like physical models, biogeochemical models generally undergo long spinup integrations to minimize drift. Nevertheless, based on a limited subset of models, it is found that drift is an important consideration and must be accounted for. For properties or regions where drift is important, the drift correction method must be carefully considered. The use of a drift estimate based on the full control time series is recommended to minimize the contamination of the drift estimate by internal variability.
    Suga T., K. Motoki, Y. Aoki, and A. M. Macdonald, 2004: The North Pacific climatology of winter mixed layer and mode waters. J. Phys. Oceanogr.,34, 3-22, doi: 10.1175/1520-0485.8d225dc68ccd5cd3f64570bb3fdcb720http%3A%2F%2Fadsabs.harvard.edu%2Fcgi-bin%2Fnph-data_query%3Fbibcode%3D2004JPO....34....3S%26db_key%3DPHY%26link_type%3DABSTRACT%26high%3D17368/s?wd=paperuri%3A%28b596333cea7ef3c37b6c621ea2dd4749%29&filter=sc_long_sign&tn=SE_xueshusource_2kduw22v&sc_vurl=http%3A%2F%2Fadsabs.harvard.edu%2Fcgi-bin%2Fnph-data_query%3Fbibcode%3D2004JPO....34....3S%26db_key%3DPHY%26link_type%3DABSTRACT%26high%3D17368&ie=utf-8
    Taylor K. E., R. J. Stouffer, and G. A. Meehl, 2012: An overview of CMIP5 and the experiment design. Bull. Amer. Meteor. Soc.,93, 485-498, doi: 10.1175/BAMS-D-11-00094.1.10.1175/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.
    Toyoda T., T. Awaji, Y. Ishikawa, and T. Nakamura, 2004: Preconditioning of winter mixed layer in the formation of North Pacific eastern subtropical mode water. Geophys. Res. Lett., 31,L17206, doi: 10.1029/2004GL020677.10.1029/2004GL020677f3aa24ca7b49f966d7df8ea843ca03e7http%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1029%2F2004GL020677%2Fabstracthttp://onlinelibrary.wiley.com/doi/10.1029/2004GL020677/abstractABSTRACT The preconditioning of winter mixed layer in the formation of the eastern subtropical mode water (ESTMW) in the North Pacific Ocean is investigated using an eddy-permitting ocean general circulation model. The result shows that convergence in the northward Ekman transport of saltier water and weak summertime heat fluxes due to the presence of stratus cloud in the ESTMW formation region are both central to the initiation of significant mixed layer deepening that subsequently evolves into the ESTMW pycnostad in the main thermocline. These two factors originate from air-sea interactions inherent in the region of Northeast Pacific Basin. The approach described here has significant potential for determining the water-mass distribution and may lead to an explanation of the Sverdrup's Eastern Gyral in the subtropical North Pacific.
    Toyoda T., T. Awaji, S. Masuda, N. Sugiura, H. Igarashi, T. Mochizuki, and Y. Ishikawa, 2011: Interannual variability of North Pacific eastern subtropical mode water formation in the 1990s derived from a 4-dimensional variational ocean data assimilation experiment. Dyn. Atmos.Oceans, 51, 1- 25.
    Wang L. Y., Q. Y. Liu, L. X. Xu, and S.-P. Xie, 2013: Response of mode water and subtropical countercurrent to greenhouse gas and aerosol forcing in the North Pacific. Journal of Ocean University of China,12, 222-229, doi: 10.1007/s11802-013-2193-x.10.1007/s11802-013-2193-x4fef76ef-857d-4edf-a5b3-988328a897b24709b3da7a2372f728009988e4429ba0http%3A%2F%2Fwww.cnki.com.cn%2FArticle%2FCJFDTotal-QDHB201302005.htmrefpaperuri:(db09f2a064623c26259f6245b12f5ad6)http://d.wanfangdata.com.cn/Periodical_qdhydxxb-e201302005.aspxThe response of the North Pacific Subtropical Mode Water and Subtropical Countercurrent (STCC) to changes in greenhouse gas (GHG) and aerosol is investigated based on the 20th-century historical and single-forcing simulations with the Geo-physical Fluid Dynamics Laboratory Climate Model version 3 (GFDL CM3). The aerosol effect causes sea surface temperature (SST) to decrease in the mid-latitude North Pacific, especially in the Kuroshio Extension region, during the past five decades (1950-2005), and this cooling effect exceeds the warming effect by the GHG increase. The STCC response to the GHG and aerosol forcing are opposite. In the GHG (aerosol) forcing run, the STCC decelerates (accelerates) due to the decreased (increased) mode waters in the North Pacific, resulting from a weaker (stronger) front in the mixed layer depth and decreased (increased) subduction in the mode water formation region. The aerosol effect on the SST, mode waters and STCC more than offsets the GHG effect. The response of SST in a zonal band around 40?N and the STCC to the combined forcing in the historical simulation is similar to the response to the aerosol forcing.
    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, 966- 986.10.1175/2009JCLI3329.179270f55-31c4-4e86-9ac0-3fc7126cf80d3eefc59c87e4f4ca7ffcbe050a36a436http%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, 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.
  • [1] WANG Mingxing, LIU Qiang, YANG Xin, 2004: A Review of Research on Human Activity Induced Climate Change I. Greenhouse Gases and Aerosols, ADVANCES IN ATMOSPHERIC SCIENCES, 21, 314-321.  doi: 10.1007/BF02915561
    [2] YUE Yanyu, NIU Shengjie, ZHAO Lijuan, ZHANG Yu, XU Feng, 2014: The Influences of Macro- and Microphysical Characteristics of Sea-Fog on Fog-Water Chemical Composition, ADVANCES IN ATMOSPHERIC SCIENCES, 31, 624-636.  doi: 10.1007/s00376-013-3059-2
    [3] Israel LOPEZ-COTO, Subhomoy GHOSH, Kuldeep PRASAD, James WHETSTONE, 2017: Tower-Based Greenhouse Gas Measurement Network Design——The National Institute of Standards and Technology North East Corridor Testbed, ADVANCES IN ATMOSPHERIC SCIENCES, 34, 1095-1105.  doi: 10.1007/s00376-017-6094-6
    [4] FAN Xuehua, XIA Xiang'ao, CHEN Hongbin, 2015: Comparison of Column-Integrated Aerosol Optical and Physical Properties in an Urban and Suburban Site on the North China Plain, ADVANCES IN ATMOSPHERIC SCIENCES, 32, 477-486.  doi: 10.1007/s00376-014-4097-0
    [5] WANG Dongxiao, WANG Jia, Lixin WU, Zhengyu LIU, 2003: Regime Shifts in the North Pacific Simulated by a COADS-driven Isopycnal Model, ADVANCES IN ATMOSPHERIC SCIENCES, 20, 743-754.  doi: 10.1007/BF02915399
    [6] Wu Renguang, Chen Lieting, 1995: Interannual Fluctuations of Surface Air Temperature over North America and Its Relationship to the North Pacific SST Anomaly, ADVANCES IN ATMOSPHERIC SCIENCES, 12, 20-28.  doi: 10.1007/BF02661284
    [7] Guanghui ZHOU, Rong-Hua ZHANG, 2022: Structure and Evolution of Decadal Spiciness Variability in the North Pacific during 2004–20, Revealed from Argo Observations, ADVANCES IN ATMOSPHERIC SCIENCES, 39, 953-966.  doi: 10.1007/s00376-021-1358-6
    [8] LIU Qinyu, WEN Na, YU Yongqiang, 2006: The Role of the Kuroshio in the Winter North Pacific Ocean-Atmosphere Interaction: Comparison of a Coupled Model and Observations, ADVANCES IN ATMOSPHERIC SCIENCES, 23, 181-189.  doi: 10.1007/s00376-006-0181-4
    [9] LI Chongyin, XIAN Peng, 2003: Atmospheric Anomalies Related to Interdecadal Variability of SST in the North Pacific, ADVANCES IN ATMOSPHERIC SCIENCES, 20, 859-874.  doi: 10.1007/BF02915510
    [10] M. Mohsin IQBAL, M. Arif GOHEER, 2008: Greenhouse Gas Emissions from Agro-Ecosystems and Their Contribution to Environmental Change in the Indus Basin of Pakistan, ADVANCES IN ATMOSPHERIC SCIENCES, 25, 1043-1052.  doi: 10.1007/s00376-008-1043-z
    [11] A. Longhetto, S. Ferrarese, C. Cassardo, C. Giraud, F. Apadttla, P. Bacci, P. Bonelli, A. Marzorati, 1997: Relationships between Atmospheric Circulation Patterns and CO2 Greenhouse-Gas Concentration Levels in the Alpine Troposphere, ADVANCES IN ATMOSPHERIC SCIENCES, 14, 309-322.  doi: 10.1007/s00376-997-0052-7
    [12] Yao HA, Zhong ZHONG, Haikun ZHAO, Yimin ZHU, Yao YAO, Yijia HU, 2022: A Climatological Perspective on Extratropical Synoptic-Scale Transient Eddy Activity Response to Western Pacific Tropical Cyclones, ADVANCES IN ATMOSPHERIC SCIENCES, 39, 333-343.  doi: 10.1007/s00376-021-0375-9
    [13] Sandeep D. WAGH, Baban NAGARE, Sanjay D. MORE, P. Pradeep KUMAR, 2017: Multiyear Observations of Deposition-Mode Ice Nucleating Particles at Two High-Altitude Stations in India, ADVANCES IN ATMOSPHERIC SCIENCES, 34, 1437-1446.  doi: 10.1007/s00376-017-7048-8
    [14] Wang Mingxing, Zhang Renjian, Pu Yifen, 2001: Recent Researches on Aerosol in China, ADVANCES IN ATMOSPHERIC SCIENCES, 18, 576-586.  doi: 10.1007/s00376-001-0046-9
    [15] Chunsheng ZHAO, Yingli YU, Ye KUANG, Jiangchuan TAO, Gang ZHAO, 2019: Recent Progress of Aerosol Light-scattering Enhancement Factor Studies in China, ADVANCES IN ATMOSPHERIC SCIENCES, 36, 1015-1026.  doi: 10.1007/s00376-019-8248-1
    [16] HUO Juan, LU Daren, 2010: Preliminary Retrieval of Aerosol Optical Depth from All-sky Images, ADVANCES IN ATMOSPHERIC SCIENCES, 27, 421-426.  doi: 10.1007/s00376-009-8216-2
    [17] Xingmin LI, Yan DONG, Zipeng DONG, Chuanli DU, Chuang CHEN, 2016: Observed Changes in Aerosol Physical and Optical Properties before and after Precipitation Events, ADVANCES IN ATMOSPHERIC SCIENCES, 33, 931-944.  doi: 10.1007/s00376-016-5178-z
    [18] Yuhuan LÜ, Hengchi LEI, Jiefan YANG, 2017: Aircraft Measurements of Cloud-Aerosol Interaction over East Inner Mongolia, ADVANCES IN ATMOSPHERIC SCIENCES, 34, 983-992.  doi: 10.1007/s00376-017-6242-z
    [19] Xiaoyan WU, Jinyuan XIN, Wenyu ZHANG, Chongshui GONG, Yining MA, Yongjing MA, Tianxue WEN, Zirui LIU, Shili TIAN, Yuesi WANG, Fangkun WU, 2020: Optical, Radiative and Chemical Characteristics of Aerosol in Changsha City, Central China, ADVANCES IN ATMOSPHERIC SCIENCES, 37, 1310-1322.  doi: 10.1007/s00376-020-0076-9
    [20] WANG Zhili, ZHANG Hua, SHEN Xueshun, Sunling GONG, ZHANG Xiaoye, 2010: Modeling Study of Aerosol Indirect Effects on Global Climate with an AGCM, ADVANCES IN ATMOSPHERIC SCIENCES, 27, 1064-1077.  doi: 10.1007/s00376-010-9120-5

Get Citation+

Export:  

Share Article

Manuscript History

Manuscript received: 13 April 2015
Manuscript revised: 25 September 2015
Manuscript accepted: 05 November 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

Response of North Pacific Eastern Subtropical Mode Water to Greenhouse Gas Versus Aerosol Forcing

  • 1. Physical Oceanography Laboratory, Ocean University of China, Qingdao 266100

Abstract: Mode water is a distinct water mass characterized by a near vertical homogeneous layer or low potential vorticity, and is considered essential for understanding ocean climate variability. Based on the output of GFDL CM3, this study investigates the response of eastern subtropical mode water (ESTMW) in the North Pacific to two different single forcings: greenhouse gases (GHGs) and aerosol. Under GHG forcing, ESTMW is produced on lighter isopycnal surfaces and is decreased in volume. Under aerosol forcing, in sharp contrast, it is produced on denser isopycnal surfaces and is increased in volume. The main reason for the opposite response is because surface ocean-to-atmosphere latent heat flux change over the ESTMW formation region shoals the mixed layer and thus weakens the lateral induction under GHG forcing, but deepens the mixed layer and thus strengthens the lateral induction under aerosol forcing. In addition, local wind changes are also favorable to the opposite response of ESTMW production to GHG versus aerosol.

1. Introduction
  • Mode water is a distinct water mass with vertical consistency of temperature, density and salinity, found in the permanent thermocline. The main feature of mode water is that it can carry low potential vorticity (PV) water from the mixed layer into the subsurface interior, which makes the upper ocean and the underlying layers connected. Formation regions of mode waters are of particular importance for issues of climate change, since they are areas where atmospheric variability is transmitted to large volumes of water that then penetrate, and are furthermore sequestered, below the surface layer. These regions therefore provide a link between the shorter time scales of the atmosphere and the longer time scales of the oceanic subsurface layers. Mode waters that are subducted in the northern region of the subtropical gyre impact the water-mass properties of the thermocline, and they are characterized by recirculation patterns with decadal time scales. Indeed, variability in these mode waters has been hypothesized to be an important part of midlatitude ocean-atmosphere feedback scenarios of decadal variability (e.g., Latif and Barnett, 1994). Variations in their properties, including temperature, salinity and PV, could impact decadal feedback directly by altering ocean temperature, and indirectly by altering the circulation.

    The subduction of mode waters can be accomplished by Ekman pumping and lateral induction (Marshall et al., 1993), and the latter has been found to play a more important role in the formation of mode waters (Huang and Qiu, 1994; Nakamura, 1996). This is in agreement with the work of (Inui et al., 1999), who found that the path of low-PV water (i.e., mode water) on each isopycnal surface starts from points where the isopycnal outcropping line intersects the winter mixed layer depth (MLD) horizontal gradients (i.e., MLD front). The overlap of deep mixed layer regions with isopycnal outcrops is critical in this process; the deep mixed layers provide the low PV water, whereas the layer outcrop sets the density and provides the pathways into the interior. Once the mode water has been isolated from surface forcing, horizontal advection spreads it and its evolution is well beyond the formation area.

    In the North Pacific, three kinds of mode water have been identified. Western subtropical mode water (WSTMW) is formed to the south of the Kuroshio Extension and is distributed around the northwestern part of the subtropical gyre, and central mode water (CMW) is formed to the north of the Kuroshio Extension and is distributed in the central part of the northern subtropical gyre. The third one, eastern subtropical mode water (ESTMW), is formed in the eastern part of the subtropical gyre. ESTMW is centered at (30°N, 140°W) with a temperature range of 16°C-22°C and a density range of 24.0-25.4 kg m-3 (Hautala and Roemmich, 1998). Despite it being weaker than WSTMW and CMW in terms of thickness and area of extent (Hanawa and Talley, 2001), ESTMW is worthy of study because its formation region sits along an important conduit for the subtropical-tropical inter-gyre exchange in the North Pacific Ocean.

    Recent studies have investigated the formation mechanism as well as the climate variability of ESTMW. For example, (Toyoda et al., 2004) studied its climatological formation mechanism and demonstrated the importance of the preconditioning of the mixed layer over the formation region. Further, (Toyoda et al., 2011) examined the interannual variability of ESTMW by analyzing a 1990s ocean dataset from a 4D variational data assimilation experiment. (Hu et al., 2011) investigated the mechanism of interannual-to-decadal variability of the subduction rate over the ESTMW formation area during 1958-2001 and found the variability of the wind stress curl plays a dominant role. To explore its response to global warming, (Luo et al., 2009) compared solutions from a set of coupled climate models between a present-day climate and a future warmer climate and found that global warming induces a significant reduction in the volume of the ESTMW. This change is due to a more stratified upper ocean and a shoaling of the winter MLD, resulting mainly from a reduction of the ocean-to-atmosphere heat loss over the eastern subtropical North Pacific.

    Of the human-caused changes to climate forcing, while greenhouse gases (GHGs) warm the planet, aerosols (tiny suspended particles or droplet in the atmosphere) tend to cool. Aerosols differ from GHGs in that their lifetime in the atmosphere is short (a few weeks), before they settle or wash out. Therefore, the cooling effect from aerosols is believed to be regional rather than global, like it is for GHGs (Mitchell et al., 1995). Based on GFDL CM3 output, (Wang et al., 2013) recently evaluated both the warming effect of GHGs and the cooling effect of aerosols on WSTMW, as well as a closely related subtropical countercurrent in the North Pacific. Their results show that GHG (aerosol) forcing induces a weakening (strengthening) of WSTMW and thus a deceleration (acceleration) of the subtropical countercurrent, and the cooling effect of aerosols exceeds the warming effect of GHGs.

    For this study, we pay particular attention to ESTMW and examine how the aerosol-induced cooling effect competed with the GHG-induced warming effect in the formation of ESTMW in the North Pacific during the past century. Our main finding is that the response of ESTMW to GHGs and aerosols is just the opposite. Under GHG forcing, ESTMW is decreased in volume and produced on lighter isopycnal surfaces, while it is produced on denser isopycnal surfaces and its volume is significantly increased under aerosol forcing.

    This paper is organized as follows: Section 2 briefly describes the model and simulations. Section 3 examines the response of ESTMW in the North Pacific to GHG versus aerosol forcing. Section 4 discusses the formation mechanisms of the opposite response of ESTMW to GHG and aerosol, and a summary is given in section 5.

2. Model and data
  • We analyze simulations from the newest version of NOAA-GFDL's climate model, CM3. Compared with its earlier version, CM2.1, CM3 represents an evolutionary step with emphasis on improving formulational aspects and capabilities of the atmosphere and land components (Donner et al., 2011; Griffies et al., 2011). CM3 uses a flexible modeling system to couple the GFDL Atmospheric Model version 3 (AM3) with the Modular Ocean Model version 4p1 (MOM4p1). AM3 employs a cubed-sphere implementation of the finite-volume dynamical core. Its horizontal resolution is approximately 200 km, and its vertical resolution ranges from approximately 70 m near the Earth's surface to 1-1.5 km near the tropopause and 3-4 km in much of the stratosphere. MOM4p1 employs a finite difference approach to solve the primitive equations. Its resolution is 1° in both latitude and longitude, with refined meridional resolution equatorward of 30° so that it reaches (1/3)° at the equator. The model has 50 levels in the vertical direction with 22 levels of 10-m thickness in the top 220 m. A tripolar horizontal grid with poles over Eurasia, North America and Antarctica is used to avoid polar filtering over the Arctic. Through experiments with historical radiative forcing, notable improvements are apparent in the CM3 climate simulation relative to CM2.1, including more realistic aerosol direct effects, a better shortwave cloud forcing, a modified SST bias pattern, and reduced biases in the Arctic SLP and sea-ice cover.

    As a major U.S. climate model contributing to CMIP5, a number of experiments with CM3 have been performed following CMIP5 protocol (Taylor et al., 2012). These runs include pre-industrial control, historical all forcing, as well as single forcing simulations for the years 1860-2005. The historical all forcing runs enable evaluation of model performance against present climate and observed climate change, and the single forcing runs aim to attribute observed climate change to particular causes, in which two major experiments are performed to yield estimates of the contribution of GHG forcing (GHG-only simulation) and aerosol forcing (aerosol-only simulation) to climate change.

    In this paper, we analyze solutions from the GHG-only, aerosol-only, and historical all forcing simulations. Three members of each simulation are selected and their ensemble means are obtained, which effectively remove natural variability from an individual simulation. The results presented below are through two types of analyses: trends are the linear least-squares regressions of each quantity over 1900-2005, and monthly climatology is averaged over 1950-2005. A note is that the pre-industrial control experiment is used as a reference, and the trend analysis is performed on the difference of the ensemble runs from the control run.

    Figure 1.  Mean depth (color shading; units: m) and density (contour interval = 0.2 kg m$^-3$) of the March mixed layer from the (a) GHG-only simulation, (b) aerosol-only simulation, and (c) historical all forcing simulation. Trends of March mixed layer depth [color shading; units: m (100 yr)$^-1$] from the (d) GHG-only simulation, (e) aerosol-only simulation, and (f) historical all forcing simulation. In this and following figures, trends are the linear least-squares regressions of each quantity over 1900-2005.

3. Results
  • As mentioned previously the lateral induction process is more important in producing ESTMW in the North Pacific. Since the lateral induction, defined as horizontal advection across the sloping mixed layer base, becomes large during winter when the mixed layer deepens, in this section we examine both the winter MLD and lateral induction.

    The MLD reaches its seasonal maximum during winter in the North Pacific. For the historical run (Fig. 1c), a local maximum (200 m) is found in the eastern subtropics centered at (30°N, 140°W), corresponding to the formation region of ESTMW (Hosoda et al., 2001). While the maximum winter MLD appears to be deeper in the model, its spatial pattern is in good agreement with observations (Suga et al., 2004; Toyoda et al., 2011).

    Comparing the GHG-only run (Fig. 1a) with the historical all forcing run (Fig. 1c), although the overall horizontal spatial structure of the winter MLD does not change much, it shoals in the eastern subtropics with its maximum approximately 40 m shallower near the center. Such a change in the mixed layer is associated with smaller MLD horizontal gradients, which in turn will result in weaker lateral induction and thus a decrease of ESTMW. Comparing the aerosol-only run (Fig. 1b) with the historical all forcing run (Fig. 1c), however, the local mixed layer in the eastern subtropics is seen to deepen, with its maximum approximately 20 m deeper around the center. This change in the mixed layer is associated with larger MLD horizontal gradients, which in turn will result in stronger lateral induction and thus an increase of ESTMW. Furthermore, in comparison with the historical all forcing run, the mixed layer density appears to be lighter in the GHG-only run but denser in the aerosol-only forcing run.

    Over the eastern subtropical region of the North Pacific, therefore, we find that, in stark contrast to the GHG forcing, which produces a shallower mixed layer and thus a weaker MLD front, the aerosol forcing induces a deeper mixed layer and thus a stronger MLD front. The opposite response of the MLD to the GHG and aerosol forcing can be further verified by their trends. For the GHG run (Fig. 1d), the MLD exhibits a falling trend over the entire eastern basin, with more significance close to the ESTMW formulation center where the maximum trend exceeds -60 m (100 yr)-1. This indicates that GHG forcing induces more shoaling in the maximum MLD area compared to its surrounding regions, resulting in a weaker MLD front and thus a decrease of lateral induction. For the aerosol run (Fig. 1e), just the opposite occurs; the MLD exhibits a distinct rising trend [60 m (100 yr)-1] in the deep MLD area but a falling trend [-40 m (100 yr)-1] to the southeast. Such a change in the MLD leads to a stronger MLD front and thus an increase of lateral induction. For the historical run (Fig. 1f), as a result of the combination of the two forcings, the spatial pattern of the MLD trend is similar to that in the aerosol run, indicating that the aerosol effect exceeds the GHG effect on the MLD.

    Figure 2.  Mean lateral induction (color shading; units: m yr$^-1$) in March from the (a) GHG-only simulation, (b) aerosol-only simulation, and (c) historical all forcing simulation. Trends of March lateral induction [color shading; units: m yr$^-1$ (100 yr)$^-1$] from the (d) GHG-only simulation, (e) aerosol-only simulation, and (f) historical all forcing simulation.

    Next we examine the lateral induction, which is believed to play an important role in producing the ESTMW in the North Pacific. The lateral induction is estimated by u m· ▽ h m, in which u m is the horizontal velocity at the bottom of the mixed layer and h m is the MLD. Figure 2 compares the response of the lateral induction to GHG and aerosol forcing with the historical run. In the eastern part of the subtropical gyre, large lateral induction appears to the southeast of the deep MLD where the strong MLD front is located. For the historical run (Fig. 2c), its maximum exceeds 130 m yr-1. While the area of the lateral induction is found to shrink due to GHG forcing, with its maximum being reduced to only about 100 m yr-1 (Fig. 2a), the area expends largely due to aerosol forcing, with its maximum being enhanced to 160 m yr-1 (Fig. 2b). The opposite response of the lateral induction is further confirmed by their linear trends, i.e., to the southeast of the deep MLD area the GHG forcing induces a falling trend of -60 m yr-1 (100 yr)-1 (Fig. 2d) but the aerosol forcing induces a rising trend of 100 m yr-1 (100 yr)-1 (Fig. 2e). In addition, the result of the historical run suggests that the cooling effect of aerosols exceeds the warming effect of GHGs on the lateral induction (Fig. 2f), consistent with the findings above from the response of the MLD.

    Figure 3.  The (a) mean and (b) trend of May low-PV ($<1.5\times10^-10$ m$^-1$ s$^-1$) water volume for each density class within the region between 130$^\circ$-160$^\circ$W and 15$^\circ$-25$^\circ$N.

    Figure 4.  Potential vorticity (color shading; units: 10$^-11$ m$^-1$ s$^-1$) in May in the core density classes from the (a) GHG-only simulation, (b) aerosol-only simulation, and (c) historical all forcing simulation. Potential vorticity (color shading; units: 10$^-11$ m$^-1$ s$^-1$) and density (contour interval = 0.2 kg m$^-3$) in May along the 20.5$^\circ$N section from the (d) GHG-only simulation, (e) aerosol-only simulation, and (f) historical all forcing simulation. Potential vorticity (color shading; units: 10$^-11$ m$^-1$ s$^-1$) and density (contour interval = 0.2 kg m$^-3$) in May along the 150$^\circ$W section from the (g) GHG-only simulation, (h) aerosol-only simulation, and (i) historical all forcing simulation.

    Figure 5.  Ocean-to-atmosphere heat flux changes (color shading; units: W m$^-2$) from the pre-industrial control experiment in the (a) GHG-only simulation, (b) aerosol-only simulation, and (c) historical all forcing simulation. Superimposed is the climatological ocean-to-atmosphere heat flux (contour interval = 10 W m$^-2$) in the control experiment.

    Figure 6.  Trends of horizontal velocity [color shading; units: cm s$^-1$ (100 yr)$^-1$] along the slanted section (thick red line in Fig. 1a) from the (a) GHG-only simulation, (b) aerosol-only simulation, and (c) historical all forcing simulation. Superimposed are the climatological velocity (contour interval = 1 cm s$^-1$) and mixed layer depth (circled line) in each experiment.

  • Mode waters are characterized by the vertical homogeneity of their properties, which is quantitatively summarized by its PV. The PV is defined as (f/ρ)(∆ρ/∆ z), in which f is the Coriolis parameter, ρ is seawater density, and ∆ρ/∆ z is the vertical gradient of the density. After strong subduction during the winter season, the intensity of newly formed mode water can be measured by low-PV water volume in late spring near the formation region. Figure 3a shows the volume of the low-PV water in May for each density class from 24.6 to 25.8 kg m-3 between the layers defined by the adjacent isopycnals at 0.1 kg m-3 interval within the region between 130°-160°W and 15°-25°N.

    If the ESTMW is represented by PV<1.5× 10-10m-1 s-1, the core density classes are found to be 25.0-25.2 kg m-3 in the GHG run, 25.4-25.6 kg m-3 in the aerosol run, and 25.3-25.5 kg m-3 in the historical, respectively, i.e., the ESTMW forms on lighter isopycnal surfaces due to GHG forcing but on denser isopycnal surfaces due to aerosol forcing. In addition, the ESTMW volume in the core density layers appears to be smallest in the GHG forcing run, largest in the aerosol forcing run, and somewhere in between in the historical all forcing run. To be more specific (Table 1), its volume is increased by 0.48× 1014 m3, or 17%, from 2.80× 1014 m3 in the historical simulation to 3.28× 1014 m3 in the aerosol-only simulation, while it is decreased by 1.46× 1014 m3, or 52%, to 1.34× 1014 m3 in the GHG-only simulation. Figure 3b shows the linear trends of the low-PV water volume for each density class in the scenario runs. It is clear that the volume trend in the GHG run is falling for all density classes, with most significance from 25.2 to 25.4 kg m-3, while the volume trend in the aerosol run is rising for all density classes, with most significance from 25.3 to 25.5 kg m-3. Additionally, the aerosol effect appears again to exceed the GHG effect, resulting in a rising trend of the ESTMW volume in the historical run. Note there is a trend in the control run (Fig. 3b),which is the climate drift. In the absence of external forcing, climate models often exhibit long-term trends that cannot be attributed to natural variability. This so-called climate drift arises for various reasons, including perturbations to the climate system in the coupling together of component models, and deficiencies in model physics and numerics. Climate drift is common among the CMIP5 models (Sen Gupta et al., 2013).

    The distributions of May PV are shown in Fig. 4 on the core density isopycnal surfaces, as well as along latitudinal and longitudinal sections for the three simulations. Comparing the aerosol-only simulation with the historical simulation, it can be clearly seen that the region with low-PV water in the eastern subtropics is expanded both horizontally and vertically, indicating that the aerosol forcing induces a significant increase in volume of the ESTMW. Comparing the GHG-only simulation with the historical simulation, we see that the region with low-PV water is reduced both horizontally and vertically, indicating that the GHG forcing induces a reduction in volume of the ESTMW, which is in agreement with previous studies (Luo et al., 2009).

    Figure 7.  Wind stress changes from the pre-industrial control experiment in the (a) GHG-only simulation, (b) aerosol-only simulation, and (c) historical all forcing simulation.

    Figure 8.  Ekman pumping changes (color shading; units: 10$^-8$ m s$^-1$) from the pre-industrial control experiment in the (a) GHG-only simulation, (b) aerosol-only simulation, and (c) historical all forcing simulation. Superimposed is the climatological Ekman pumping (contour interval = $0.4\times 10^-8$ m s$^-1$) in the control experiment.

    Figure 9.  Trends of temperature [color shading; units: $^\circ$C (100 yr)$^-1$] along the slanted section (thick red line in Fig. 1a) from the (a) GHG-only simulation, (b) aerosol-only simulation, and (c) historical all forcing simulation. Superimposed is the climatological temperature (contour interval = 2$^\circ$C) and mixed layer depth (circled line) in each experiment.

4. Formation mechanisms
  • As demonstrated above, a spatially non-uniform shoaling (deepening) of the mixed layer in the eastern subtropics is associated with a weakening (strengthening) of lateral induction, corresponding to the decreased (increased) ESTMW in the North Pacific under GHG (aerosol) forcing. Since the mixed layer change is typically related to the surface forcing change, here we examine the changes in the surface heat flux and wind stress in both the GHG-only and aerosol-only simulations.

    Under GHG forcing, the ocean-to-atmosphere heat loss (Fig. 5) is reduced around the center of the ESTMW formation area but increased to its southeast. Such a change in the heat flux results in decreased MLD horizontal gradients and thus reduced lateral induction, leading to less mode water formed in the eastern subtropical gyre. Under aerosol forcing, however, the heat loss over the region is significantly increased, with the maximum being more than 12 W m-2 around the center. This change of the surface heat flux produces a less stratified upper ocean and a resultant deepening of the MLD. As a result, the lateral induction is increased and more mode water is produced in the eastern subtropics. For the historical run, similar to what happens under aerosol forcing, the heat loss is increased over the region and thus the formation of the ESTMW is intensified. Among the four physical components included in the heat flux (i.e., shortwave radiation, longwave radiation, sensible heat flux, and latent heat flux), further analysis indicates that the contribution from the latent heat loss is predominant (not shown).

    As mentioned previously, the lateral induction is associated with the MLD front and the horizontal velocity across the sloping mixed layer base. Now that changes in the former have been found to play an important role in the response of the lateral induction to GHG and aerosol forcing, next we examine if changes in the latter also contribute to the lateral induction response. To examine this, we select a slanted section across the ESTMW formation region from (38°N, 150°W) to (20°N, 128°W) (the thick red line in Fig. 1a). Figure 6 shows the horizontal velocity changes in the GHG, aerosol and historical runs along the slanted section, respectively. For the mean velocity fields (contours in Fig. 6), a southeastward current of 1 cm s-1 appears over the large MLD front area, which is apparently favorable to the formation of the ESTMW. Under GHG (aerosol) forcing, this southeastward velocity is reduced (increased) by approximately 0.2 cm s-1, contributing to the lesser (greater) production of ESTMW. For the historical run, as a result of the combination of the two forcing effects, its change is similar to that for the aerosol run, demonstrating that the aerosol effect exceeds the GHG effect on the velocity fields. Therefore, the above analysis suggests that the response of lateral induction to GHG and aerosol forcing is a result of changes in both the strength of the MLD front and the horizontal velocity across the sloping mixed layer base over the ESTMW formation region.

    To estimate the relative contribution from the changes in the MLD versus horizontal velocity to the lateral induction, we choose a small area where the changes in the lateral induction reach their maxima under both GHG and aerosol forcing. It is found that the lateral induction is reduced by 30 m yr-1 from 86 m yr-1 in the control run to 56 m yr-1 in the GHG run, with 80% of this reduction (i.e., 24 m yr-1) from the MLD change and 20% (i.e., 6 m yr-1) from the horizontal velocity change. Meanwhile, it is increased by 34 m yr-1 from 86 m yr-1 in the control run to 120 m yr-1 in the aerosol run, with 68% of this increase (i.e., 23 m yr-1) from the MLD change and 32% (i.e., 11 m yr-1) from the horizontal velocity change. Therefore, the MLD change plays a dominant role in the change of ESTMW in the North Pacific under both GHG and aerosol forcing.

    For the changes in wind stress (Fig. 7), the northeast trade winds are weakened (intensified) under GHG (aerosol) forcing, leading directly to local mixed layer shoaling (deepening) due to weaker (stronger) wind stirring. Those changes in the latent heat flux described above can be caused by the weakened (intensified) northeast trade winds under GHG (aerosol) forcing, which promotes less (more) heat loss from the ocean into the atmosphere through the wind-evaporation-SST feedback mechanism (Xie et al., 2010).

    In addition to lateral induction, the subduction of ESTMW in the North Pacific can also be accomplished through Ekman pumping (e.g., Huang and Qiu, 1994). Under GHG forcing, the wind changes result in a reduced Ekman pumping (downward is negative) in the eastern subtropical North Pacific (Fig. 8a), weakening the ESTMW formation. Under aerosol forcing, the Ekman pumping is significantly enhanced over the region (Fig. 8b), strengthening the ESTMW formation. For the historical run, similar to what happens under aerosol forcing, the Ekman pumping over the region is increased and thus the formation of the ESTMW is intensified (Fig. 8c). For the central area between 150°-130°W and 20°-35°N, the Ekman pumping is found to be 34 m yr-1 in the GHG run, 37 m yr-1 in the aerosol run, and 35 m yr-1 in the historical run.

    Figure 9 shows the vertical distribution of temperature along the slanted section for the three scenario runs. It is interesting to see that there is a cold (warm) anomaly right beneath the mixed layer around the center of the ESTMW formation region under GHG (aerosol) forcing. Such a change in the temperature in the vertical direction results from the changes in the mixed layer depth and the thermocline, due to the changes in the surface heat flux and winds in response to the GHG and aerosol forcing. A vertical movement of the thermocline changes the subsurface temperature through Ekman pumping, directly forced by the wind stress curl change, which appears to be positive (negative) under the GHG (aerosol) forcing over the eastern subtropical North Pacific (Figs. 8a and b). Note that positive (negative) wind stress curl anomalies shoal (deepen) the thermocline and produce cold upwelling (warm downwelling) in the North Hemisphere. This is in agreement with the findings of (Han et al., 2006). Through an analysis of observed data and model solutions for the second half of the last century, they found that, while increased GHGs act to warm up the upper ocean by increasing downward surface heat flux, anomalous winds in the tropics cause upward Ekman pumping velocity and shoal the thermocline, resulting in an upper-thermocline cooling. For the historical run (Fig. 9c), there a warm anomaly appears from the surface down to the thermocline, with more warming around the bottom of the mixed layer. This is a result of the combined forcing from GHGs and aerosols.

5. Summary and discussion
  • By comparing solutions between the single forcing simulations and historical all forcing simulation for the past century from GFDL CM3, this study examined the response of MLD, lateral induction, and mode water in the eastern subtropics of the North Pacific to GHG and aerosol forcing. Under GHG (aerosol) forcing, ESTMW is produced on lighter (denser) isopycnal surfaces and is significantly weakened (strengthened) in terms of its formation and evolution. The weakening (strengthening) of ESTMW is due mainly to a reduction (increase) in lateral induction, resulting from a weakening (strengthening) of both MLD horizontal gradients and horizontal velocity at the bottom of the mixed layer over the southeast of the deep MLD area. Further, the weakening (strengthening) of the MLD front arises from a reduction (increase) in local ocean-to-atmosphere latent heat loss, which tends to produce a more (less) stratified upper ocean and thus a shoaling (deepening) of the MLD over the formation region of the ESTMW in the North Pacific. In addition, reduced (enhanced) Ekman pumping resulting from GHG (aerosol) induced wind change contributes directly to a lesser (greater) production of ESTMW. As the aerosol effect exceeds the GHG effect, the spatial patterns of these above changes in the historical simulation are similar to those in the aerosol simulation.

    To validate the response of the ESTMW to the warming effect of GHGs, we also analyzed the RCP4.5 and RCP8.5 simulations with GFDL CM3. The RCP simulations are future climate projections that resemble a stronger version of the GHG simulation. Results show that the production of ESTMW in the North Pacific is significantly reduced under both the RCP4.5 and RCP8.5 simulations, which is similar to the response to the GHG forcing and thus supports the findings of this study.

    It is important to mention that, limited by the fact that our study relies on only one model, the results could be model-dependent. For example, large uncertainty has been found in the magnitude and spatial pattern of anthropogenic aerosol forcing among the CMIP5 models (e.g., Rotstayn et al., 2012). Therefore, the response of ESTMW to aerosol forcing may feature large inter-model diversity. Using a range of climate models within the CMIP5 framework, further research is needed to better understand the mechanisms and the extent to which these findings from the present study are model-dependent.

    Lastly, we would like to point out that GFDL CM3 is non-eddy-resolving, and that may be called into question for simulating the ESTMW in the North Pacific. In non-eddy-resolving models, subduction on a given isopycnal surface is limited to the cross point between the 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. A recent study found that mesoscale eddies significantly enhance the total subduction rate, helping create remarkable peaks in the volume histogram that correspond to the WSTMW and CMW in the North Pacific (Xu et al., 2014).

Reference

Catalog

    /

    DownLoad:  Full-Size Img  PowerPoint
    Return
    Return