Advanced Search
Article Contents

Tides and Turbulent Mixing in the North of Taiwan Island


doi: 10.1007/s00376-018-8098-2

  • Microstructure and hydrological profiles were collected along two cross-shelf sections from the deep slope to the shallow water in the north of Taiwan Island in the summer of 2006. While the tidal currents on the shelf were dominated by the barotropic tide with the current ellipse stretched across the shelf, significant internal tides were observed on the slope. The depth-mean turbulent kinetic energy (TKE) dissipation rate on the shelf was 10-6 W kg-1, corresponding to a diapycnal diffusivity of 10-2 m2 s-1. The depth-mean TKE dissipation rate on the slope was 1 10-7 W kg-1, with diapycnal diffusivity of 3.4 10-4 m2 s-1. The shear instability associated with internal tides largely contributed to the TKE dissipation rate on the slope from the surface to 150 m, while the enhanced turbulence on the shelf was dominated by tidal or residual current dissipations caused by friction in the thick bottom boundary layer (BBL). In the BBL, the Ekman currents associated with the northeastward Taiwan Warm Current were identified, showing a near-bottom velocity spiral, which agreed well with the analytical bottom Ekman solution.
    摘要: 本文基于2006年夏季在台湾岛东北部海域从陆坡至陆架的湍流和水文断面观测,揭示了该海域的湍流量级及相关水文特征。研究发现:陆架上潮流以正压潮为主,在地形的作用下,潮汐椭圆被压缩,而在陆坡上以显著的内潮信号为主。在陆架上,深度平均的湍动能耗散率量级为 10-6 W kg-1,对应的湍流混合率量级为10-2 m2 s-1。在陆坡上,深度平均的湍动能耗散率为10-7 W kg-1,平均湍流混合率为3.410-4 m2 s-1。陆坡上层垂向水文观测显示:由内潮引起的剪切不稳定极大的贡献于湍动能耗散率,而在陆架上增强的湍动能耗散率则主要由潮流和定常流在厚底边界层内的底摩擦作用下产生。
  • 加载中
  • Baines P. G.,1982: On internal tide generation models. Deep Sea Research Part A. Oceanographic Research Papers, 29(3), 307-338, https://doi.org/10.1016/0198-0149(82)90098-X
    Cushman-Roisin B.,V. Malačič, 1997: Bottom Ekman pumping with stress-dependent eddy viscosity. J. Phys. Oceanogr., 27, 1967-1975, https://doi.org/10.1175/1520-0485(1997)027<1967:BEPWSD>2.0.CO;2
    Egbert G. D.,R. D. Ray, 2000: Significant dissipation of tidal energy in the deep ocean inferred from satellite altimeter data. Nature, 405(6788), 775-778, https://doi.org/10.1038/35015531
    Garrett C.,E. Kunze, 2006: Internal tide generation in the deep ocean. Annual Review of Fluid Mechanics, 39(1), 57-87, https://doi.org/10.1146/annurev.fluid.39.050905.110227
    Gawarkiewicz G., Coauthors, 2011: Circulation and intrusions northeast of Taiwan: Chasing and predicting uncertainty in the cold dome. Oceanography, 24(4), 110-121, https://doi.org/10.5670/oceanog.2011.99
    Grant W. D.,O. S. Madsen, 1986: The continental-shelf bottom boundary layer. Annual Review of Fluid Mechanics, 18, 265-305, https://doi.org/10.1146/annurev.fl.18.010186.001405
    Guo X. Y.,T. Yanagi, 1998: Three-dimensional structure of tidal current in the East China Sea and the yellow sea. Journal of Oceanography, 54(6), 651-668, https://doi.org/10.1007/BF02823285
    Hsueh Y.,J. Wang, and C. S. Chern, 1992: The intrusion of the Kuroshio across the continental shelf northeast of Taiwan. J. Geophys. Res., 97(C9), 14 323-14 330, https://doi.org/10.1029/92JC01401
    Huang R. X.,2010: Ocean Circulation: Wind-Driven and Thermohaline Processes. Cambridge University Press, 806 pp.
    Isobe A.,2008: Recent advances in ocean-circulation research on the Yellow Sea and East China Sea shelves. Journal of Oceanography, 64(4), 569-584, https://doi.org/10.1007/s10872-008-0048-7
    Kerry C. G.,B. S. Powell, and G. S. Carter, 2014: The impact of subtidal circulation on internal-tide-induced mixing in the Philippine sea. J. Phys. Oceanogr., 44, 3209-3224, https://doi.org/10.1175/JPO-D-13-0249.1
    Lee J. H.,I. Lozovatsky, S. T. Jang, C. J. Jang, C. S. Hong, and H. J. S. Fernando, 2006: Episodes of nonlinear internal waves in the northern East China Sea. Geophys. Res. Lett., 33(18), https://doi.org/10.1029/2006GL027136
    Liang W. D.,T. Y. Tang, Y. J. Yang, M. T. Ko, and W. S. Chuang, 2003: Upper-ocean currents around Taiwan. Deep Sea Research Part II: Topical Studies in Oceanography, 50(6-7), 1085-1105, https://doi.org/10.1016/S0967-0645(03)00011-0
    Liu Z. Y.,H. Wei, I. D. Lozovatsky, and H. J. S. Fernando, 2009: Late summer stratification, internal waves, and turbulence in the Yellow Sea. J. Mar. Syst., 77(24), 459-472, https://doi.org/10.1016/j.jmarsys.2008.11.001
    Lozovatsky I.,Z. Y. Liu, H. Wei, and H. J. S. Fernando, 2008: Tides and mixing in the northwestern East China Sea, part II: Near-bottom turbulence. Cont. Shelf Res., 28(2), 338-350, https://doi.org/10.1016/j.csr.2007.08.007
    Lozovatsky I.,Z. Y. Liu, H. Fernand o, J. Armengol, and E. Roget, 2012: Shallow water tidal currents in close proximity to the seafloor and boundary-induced turbulence. Ocean Dynamics, 62(2), 177-191, https://doi.org/10.1007/s10236-011-0495-3
    Lozovatsky I.,J.-H. Lee, H. J. S. Fernand o, S. K. Kang, and S. U. P. Jinadasa, 2015: Turbulence in the East China Sea: The summertime stratification. J. Geophys. Res., 120(3), 1856-1871, https://doi.org/10.1002/2014JC010596
    Matsuno T.,M. Shimizu, Y. Morii, H. Nishida, and Y. Takaki, 2005: Measurements of the turbulent energy dissipation rate around the shelf break in the East China Sea. Journal of Oceanography, 61(6), 1029-1037, https://doi.org/10.1007/s10872-006-0019-9
    Moum J. N.,D. R. Caldwell, J. D. Nash, and G. D. Gunderson, 2002: Observations of boundary mixing over the continental slope. J. Phys. Oceanogr., 32, 2113-2130, https://doi.org/10.1175/1520-0485(2002)032<2113:OOBMOT>2.0.CO;2
    Munk W.,C. Wunsch, 1998: Abyssal recipes II: Energetics of tidal and wind mixing. Deep Sea Research Part I: Oceanographic Research Papers, 45(12), 1977-2010, https://doi.org/10.1016/S0967-0637(98)00070-3
    Nasmyth P. V.,1970: Oceanic turbulence. PhD dissertation, University of British Columbia, 69 pp.
    Niwa Y.,T. Hibiya, 2004: Three-dimensional numerical simulation of M2 internal tides in the East China Sea. J. Geophys. Res., 109(C4), https://doi.org/10.1029/2003JC001923
    Osborn T. R.,1980: Estimates of the local rate of vertical diffusion from dissipation measurements. J. Phys. Oceanogr., 10(1), 83-89, https://doi.org/10.1175/1520-0485(1980)010<0083:EOTLRO>2.0.CO;2
    Pedlosky J.,1987: Geophysical Fluid Dynamics. 2nd ed., Springer-Verlag, 710 pp.
    Perlin A.,J. N. Moum, J. M. Klymak, M. D. Levine, T. Boyd, and P. M. Kosro, 2005: A modified law-of-the-wall applied to oceanic bottom boundary layers. J. Geophys. Res., 110, C10S10, https://doi.org/10.1029/2004JC002310
    Perlin A.,J. N. Moum, J. M. Klymak, M. D. Levine, T. Boyd, and P. M. Kosro, 2007: Organization of stratification, turbulence, and veering in bottom Ekman layers. J. Geophys. Res., 112, C05S90, https://doi.org/10.1029/2004JC002641
    Peters H.,R. Bokhorst, 2000: Microstructure observations of turbulent mixing in a partially mixed estuary. Part I: Dissipation rate. J. Phys. Oceanogr., 30(6), 1232-1244, https://doi.org/10.1175/1520-0485(2000)030<1232:MOOTMI>2.0.CO;2
    Qiu B.,N. Imasato, 1990: A numerical study on the formation of the Kuroshio counter current and the Kuroshio branch current in the East China Sea. Cont. Shelf Res., 10(2), 165-184, https://doi.org/10.1016/0278-4343(90)90028-K
    Rudnick, D. L.,Coauthors, 2003: From tides to mixing along the Hawaiian ridge. Science, 301(5631), 355-357, https://doi.org/10.1126/science.1085837
    Shih L. H.,J. R. Koseff, G. N. Ivey, and J. H. Ferziger, 2005: Parameterization of turbulent fluxes and scales using homogeneous sheared stably stratified turbulence simulations. J. Fluid Mech., 525, 193-214, https://doi.org/10.1017/S0022112004002587
    Su J.,1998: Circulation dynamics of the China seas: North of 18°N. The Sea, A. R. Robinson and K. H. Brink, Eds., John Wiley & ons Inc., 13- 19.
    Su J. L.,W. Wang, 1987: On the sources of the Taiwan warm current from the South China Sea. Chinese Journal of Oceanology and Limnology, 5(4), 299-308, https://doi.org/10.1007/BF02843812
    Su J. L.,Y. Q. Pan, and X. S. Liang, 1994: Kuroshio intrusion and Taiwan warm current. Oceanology of China Seas, D. Zhou et al., Eds., Springer, 59- 70.
    Tang T. Y.,Y. Hsueh, Y. J. Yang, and J. C. Ma, 1999: Continental slope flow northeast of Taiwan. J. Phys. Oceanogr., 29(6), 1353-1362, https://doi.org/10.1175/1520-0485(1999)029<1353:CSFNOT>2.0.CO;2
    Trowbridge J. H.,W. R. Geyer, M. M. Bowen, and A. J. Williams III, 1999: Near-bottom turbulence measurements in a partially mixed estuary: Turbulent energy balance, velocity structure, and along-channel momentum balance. J. Phys. Oceanogr., 29(12), 3056-3072, https://doi.org/10.1175/1520-0485(1999)029<3056:NBTMIA>2.0.CO;2
    Visbeck M.,2002: Deep velocity profiling using lowered acoustic Doppler current profilers: Bottom track and inverse solutions. J. Atmos. Oceanic Technol., 19(5), 794-807, https://doi.org/10.1175/1520-0426(2002)019<0794:DVPULA>2.0.CO;2
    Waterhouse, A. F.,Coauthors, 2014: Global patterns of diapycnal mixing from measurements of the turbulent dissipation rate. J. Phys. Oceanogr., 44(7), 1854-1872, https://doi.org/10.1175/JPO-D-13-0104.1
    Wolk F.,H. Yamazaki, L. Seuront, and R. G. Lueck, 2002: A new free-fall profiler for measuring biophysical microstructure. J. Atmos. Oceanic Technol., 19(5), 780-793, https://doi.org/10.1175/1520-0426(2002)019<0780:ANFFPF>2.0.CO;2
    Wu C. R.,Y. C. Hsin, T. L. Chiang, Y. F. Lin, and I. F. Tsui, 2014: Seasonal and interannual changes of the Kuroshio intrusion onto the East China Sea shelf. J. Geophys. Res., 119(8), 5039-5051, https://doi.org/10.1002/2013JC009748
    Wunsch C.,R. Ferrari, 2004: Vertical mixing, energy, and the general circulation of the oceans. Annual Review of Fluid Mechanics, 36, 281-314, https://doi.org/10.1146/annurev.fluid.36.050802.122121
    Xia C. S.,F. L. Qiao, Y. Z. Yang, J. Ma, and Y. L. Yuan, 2006: Three-dimensional structure of the summertime circulation in the Yellow Sea from a wave-tide-circulation coupled model. J. Geophys. Res., 111, C11S03, https://doi.org/10.1029/2005JC003218
    Xie X. H.,G. Y. Chen, X. D. Shang, and W. D. Fang, 2008: Evolution of the semidiurnal (M2) internal tide on the continental slope of the northern South China Sea. Geophys. Res. Lett., 35(13), https://doi.org/10.1029/2008GL034179
    Yang J. Y.,2007: An oceanic current against the wind: How does Taiwan Island steer warm water into the East China Sea? J. Phys. Oceanogr., 37(10), 2563-2569, https://doi.org/10.1175/JPO3134.1
    Yang J. Y.,D. X. Wu, and X. P. Lin, 2008: On the dynamics of the South China Sea warm current. J. Geophys. Res., 113(C8), https://doi.org/10.1029/2007JC004427
    Yoshikawa Y.,T. Endoh, T. Matsuno, T. Wagawa, E. Tsutsumi, H. Yoshimura, and Y. Morii, 2010: Turbulent bottom Ekman boundary layer measured over a continental shelf. Geophys. Res. Lett., 37(15), L15605, https://doi.org/10.1029/2010GL044156
    Zhong Y. S.,2009: The turbulent mixing and double diffusion in Yellow Sea and East China Sea and their seasonal variation. M.S. thesis, Ocean University of China. (in Chinese with English abstract)
    Zhou L.,J. W. Tian, and X. Q. Zhang, 2005: Observation of small-scale processes over shelf break in the East China Sea. Chinese Science Bulletin, 50(24), 2885-2890, https://doi.org/10.1360/982005-269
    Zhu X. M.,G. M. Liu, 2012: Numerical study on the tidal currents, tidal energy fluxes and dissipation in the China seas. Oceanologia et Limnologia Sinica, 43( 3), 669- 677. (in Chinese with English abstract)10.1007/s11783-011-0280-z7672fe6476514ebbff85652eea1d5c2dhttp%3A%2F%2Fen.cnki.com.cn%2FArticle_en%2FCJFDTOTAL-HYFZ201203045.htmhttp://en.cnki.com.cn/Article_en/CJFDTOTAL-HYFZ201203045.htmBased on the Regional Ocean Modeling System(ROMS),a successful simulation for the tides and tidal currents has been done in the Bohai Sea,the Yellow Sea and the East China Sea(BYES).By comparing the model results with the observed data in 91 gauge stations,the numerical model results show agreement with the observation well.Model results indicate that semidiurnal tidal system is dominant in the BYES.Four semidiurnal,two diurnal independent and one degraded semidiurnal tide-amphidromic points are well simulated in the BYES,and they are all anti-clockwise.In the Bohai Sea,semidiurnal tidal currents is clockwise,but anti-clockwise for diurnal tidal currents.Both semidiurnal and diurnal tidal currents are anti-clockwise mainly in the Yellow Sea,but clockwise mainly occurs in the East China Sea and Korea Strait.Thirteen tidal current-amphidromic points for semidiurnal,ten for K1 constituent and nine for O1 constituent can be found in the area shallower than 200 meters of the BYES.The distribution of diurnal tidal currents is more complex than semi-diurnal.The tidal energy fluxes of the four major constituents transported from the Pacific to the East China Sea are 118.341GW,19.525GW,5.630GW,3.871GW,respectively.The tidal energy is dissipated more than half in the Southern Yellow Sea,30% 40% in the East China Sea,then a little less in the Northern Yellow Sea,and the least in the Bohai Sea.
  • [1] Sun-Hee SHIN, Kyung-Ja HA, 2009: Implementation of Turbulent Mixing over a Stratocumulus-Topped Boundary Layer and Its Impact in a GCM, ADVANCES IN ATMOSPHERIC SCIENCES, 26, 995-1004.  doi: 10.1007/s00376-009-8145-0
    [2] Zeng Zongyong, Ma Chengsheng, 1985: OBSERVATIONS OF TEMPERATURE MICROSTRUCTURE IN THE ATMOSPHERE, ADVANCES IN ATMOSPHERIC SCIENCES, 2, 234-242.  doi: 10.1007/BF03179755
    [3] Liu Qinyu, Qin Zenghao, 1986: DYNAMICS OF NONLINEAR BAROCLINIC EKMAN BOUNDARY LAYER, ADVANCES IN ATMOSPHERIC SCIENCES, 3, 424-431.  doi: 10.1007/BF02657932
    [4] Tan Zhemin, Wang Yuan, 2002: Wind Structure in an Intermediate Boundary Layer Model Based on Ekman Momentum Approximation, ADVANCES IN ATMOSPHERIC SCIENCES, 19, 266-278.  doi: 10.1007/s00376-002-0021-0
    [5] HU Yinqiao, ZUO Hongchao, 2003: The Influence of Convergence Movement on Turbulent Transportation in the Atmospheric Boundary Layer, ADVANCES IN ATMOSPHERIC SCIENCES, 20, 794-798.  doi: 10.1007/BF02915404
    [6] SUN Jianning, JIANG Weimei, CHEN Ziyun, YUAN Renmin, 2005: A Laboratory Study of the Turbulent Velocity Characteristics in the Convective Boundary Layer, ADVANCES IN ATMOSPHERIC SCIENCES, 22, 770-780.  doi: 10.1007/BF02918721
    [7] Liu Shikuo, Huang Wei, Rong Pingping, 1992: Effects of Turbulent Dispersion of Atmospheric Balance Motions of Planetary Boundary Layer, ADVANCES IN ATMOSPHERIC SCIENCES, 9, 147-156.  doi: 10.1007/BF02657505
    [8] Jiang Weimei, Yu Hongbin, 1994: Study on the Thermal Internal Boundary Layer and Dispersion of Air Pollutant in Coastal Area by Numerical Simulation, ADVANCES IN ATMOSPHERIC SCIENCES, 11, 285-290.  doi: 10.1007/BF02658147
    [9] LIU Huizhi, LIANG Bin, ZHU Fengrong, ZHANG Boyin, SANG Jianguo, 2003: A Laboratory Model for the Flow in Urban Street Canyons Induced by Bottom Heating?, ADVANCES IN ATMOSPHERIC SCIENCES, 20, 554-564.  doi: 10.1007/BF02915498
    [10] ZHANG Guangzhi, XU Xiangde, WANG Jizhi, 2003: A Dynamic Study of Ekman Characteristics by Using 1998 SCSMEX and TIPEX Boundary Layer Data, ADVANCES IN ATMOSPHERIC SCIENCES, 20, 349-356.  doi: 10.1007/BF02690793
    [11] Wu Rongsheng, 1985: THE INFLUENCES OF OROGRAPHY UPON THE FLOW WITHIN EKMAN BOUNDARY LAYER UNDER THE APPROXIMATION OF GEOSTROPHIC MOMENTUM, ADVANCES IN ATMOSPHERIC SCIENCES, 2, 1-7.  doi: 10.1007/BF03179731
    [12] MA Yaoming, Massimo MENENTI, Reinder FEDDES, 2010: Parameterization of Heat Fluxes at Heterogeneous Surfaces by Integrating Satellite Measurements with Surface Layer and Atmospheric Boundary Layer Observations, ADVANCES IN ATMOSPHERIC SCIENCES, 27, 328-336.  doi: 10.1007/s00376-009-9024-4
    [13] HAN Bo, LU Shihua, AO Yinhuan, 2012: Development of the Convective Boundary Layer Capping with a Thick Neutral Layer in Badanjilin: Observations and Simulations, ADVANCES IN ATMOSPHERIC SCIENCES, 29, 177-192.  doi: 10.1007/s00376-011-0207-4
    [14] Yu SHI, Qingcun ZENG, Fei HU, Weichen DING, Zhe ZHANG, Kang ZHANG, Lei LIU, 2023: Different Turbulent Regimes and Vertical Turbulence Structures of the Urban Nocturnal Stable Boundary Layer, ADVANCES IN ATMOSPHERIC SCIENCES, 40, 1089-1103.  doi: 10.1007/s00376-022-2198-8
    [15] Zhu Ping, Xu Xiaojin, Li Xingsheng, 1992: A Numerical Study of Second-Order Turbulent Moments in the Stably Stratified Nocturnal Boundary Layer, ADVANCES IN ATMOSPHERIC SCIENCES, 9, 201-212.  doi: 10.1007/BF02657510
    [16] LIU Huizhi, LIANG Bin, ZHU Fengrong, ZHANG Boyin, SANG Jianguo, 2004: Water-Tank Experiment on the Thermal Circulation Induced by the Bottom Heating in an Asymmetric Valley, ADVANCES IN ATMOSPHERIC SCIENCES, 21, 536-546.  doi: 10.1007/BF02915721
    [17] Jae-Jin KIM, Jong-Jin BAIK, 2005: Physical Experiments to Investigate the Effects of Street Bottom Heating and Inflow Turbulence on Urban Street-Canyon Flow, ADVANCES IN ATMOSPHERIC SCIENCES, 22, 230-237.  doi: 10.1007/BF02918512
    [18] Jae-Jin KIM, Jong-Jin BAIK, 2010: Effects of Street-Bottom and Building-Roof Heating on Flow in Three-Dimensional Street Canyons, ADVANCES IN ATMOSPHERIC SCIENCES, 27, 513-527.  doi: 10.1007/s00376-009-9095-2
    [19] Liu Shikuo, Peng Weihong, Huang Feng, Chi Dongyan, 2002: Effects of Turbulent Dispersion on the Wind Speed Profile in the Surface Layer, ADVANCES IN ATMOSPHERIC SCIENCES, 19, 794-806.  doi: 10.1007/s00376-002-0045-5
    [20] Zeng Zongyong, Ma Chengsheng, Liu Xiaochun, Ling Huiqin, 1989: An Analysis of the Turbulent Structure in the Unstable Surface Layer nearby a Shelter Belt, ADVANCES IN ATMOSPHERIC SCIENCES, 6, 493-500.  doi: 10.1007/BF02659083

Get Citation+

Export:  

Share Article

Manuscript History

Manuscript received: 26 April 2018
Manuscript revised: 10 September 2018
Manuscript accepted: 12 October 2018
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

Tides and Turbulent Mixing in the North of Taiwan Island

    Corresponding author: Xiangzhou SONG, song@ouc.edu.cn
  • 1. Physical Oceanography Laboratory, Ocean University of China, Qingdao 266003, China
  • 2. National Marine Environmental Forecasting Center, State Oceanic Administration, Beijing 100081, China
  • 3. Horn Point Laboratory, University of Maryland Center for Environmental Science, Cambridge, Maryland 21613, U.S.A

Abstract: Microstructure and hydrological profiles were collected along two cross-shelf sections from the deep slope to the shallow water in the north of Taiwan Island in the summer of 2006. While the tidal currents on the shelf were dominated by the barotropic tide with the current ellipse stretched across the shelf, significant internal tides were observed on the slope. The depth-mean turbulent kinetic energy (TKE) dissipation rate on the shelf was 10-6 W kg-1, corresponding to a diapycnal diffusivity of 10-2 m2 s-1. The depth-mean TKE dissipation rate on the slope was 1 10-7 W kg-1, with diapycnal diffusivity of 3.4 10-4 m2 s-1. The shear instability associated with internal tides largely contributed to the TKE dissipation rate on the slope from the surface to 150 m, while the enhanced turbulence on the shelf was dominated by tidal or residual current dissipations caused by friction in the thick bottom boundary layer (BBL). In the BBL, the Ekman currents associated with the northeastward Taiwan Warm Current were identified, showing a near-bottom velocity spiral, which agreed well with the analytical bottom Ekman solution.

摘要: 本文基于2006年夏季在台湾岛东北部海域从陆坡至陆架的湍流和水文断面观测,揭示了该海域的湍流量级及相关水文特征。研究发现:陆架上潮流以正压潮为主,在地形的作用下,潮汐椭圆被压缩,而在陆坡上以显著的内潮信号为主。在陆架上,深度平均的湍动能耗散率量级为 10-6 W kg-1,对应的湍流混合率量级为10-2 m2 s-1。在陆坡上,深度平均的湍动能耗散率为10-7 W kg-1,平均湍流混合率为3.410-4 m2 s-1。陆坡上层垂向水文观测显示:由内潮引起的剪切不稳定极大的贡献于湍动能耗散率,而在陆架上增强的湍动能耗散率则主要由潮流和定常流在厚底边界层内的底摩擦作用下产生。

1. Introduction
  • Tides are one of the major energy sources for ocean mixing (Munk and Wunsch, 1998; Egbert and Ray, 2000). In the open ocean, turbulent mixing associated with internal tides (internal waves with tidal frequencies) generated by barotropic tidal flows over abrupt topography has been suggested to play an important role in sustaining meridional overturning circulation (Munk and Wunsch, 1998; Wunsch and Ferrari, 2004; Garrett and Kunze, 2006). In the coastal ocean (e.g., shallow shelves and estuaries), bottom boundary mixing associated with strong tidal currents is fundamental to drive the local circulation (Grant and Madsen, 1986; Moum et al., 2002; Xia et al., 2006). Thus, the mixing in the interior ocean is often envisioned as an energy cascade from the surface tides to turbulence either by internal tidal dissipation or bottom friction (Rudnick et al., 2003; Waterhouse et al., 2014). The East China Sea (ECS) is located in the Northwest Pacific (Fig. 1a), where strong tidal motions and large internal tides have been observed and modeled (Baines, 1982; Guo and Yanagi, 1998; Niwa and Hibiya, 2004; Garrett and Kunze, 2006). In the north of Taiwan Island (NTI), the Kuroshio on the slope intrudes onto the continental shelf (Qiu and Imasato, 1990; Hsueh et al., 1992; Tang et al., 1999; Liang et al., 2003; Isobe, 2008; Gawarkiewicz et al., 2011; Wu et al., 2014). On the shelf, the Taiwan Warm Current (TWC) is another dominant current (Su, 1998; Yang, 2007; Isobe, 2008). Understanding the interaction and exchange between these current systems requires a clear picture of the local turbulent mixing.

    Figure 1.  (a) Map of the major currents in the northwestern Pacific (NWP), ECS and South China Sea (SCS). The TWC and Kuroshio Current (KC) are indicated by the red arrow and curve (solid thick line). The blue line is the 200-m contour of topography from the General Bathymetric Chart of the Oceans (GEBCO). The black frame in (a) is the study area shown in (b). The AVISO climatological mean (1993-2012) absolute dynamic topography contours of 0.9 m, 1.0 m and 1.1 m are indicated by solid black lines. The colored background is the bathymetry from GEBCO. The interval of the topography is 100 m. The magenta dots represent the hydrological stations. A02 and A06 with magenta squares are 25-h repeated stations. The microstructure stations are denoted by triangles.

    Observations in the northwestern part of the ECS——namely, the Yellow Sea (Lozovatsky et al., 2008; Liu et al., 2009), northern ECS (Lee et al., 2006; Lozovatsky et al., 2012, Lozovatsky et al., 2015), and shelf break (Matsuno et al., 2005; Zhou et al., 2005)——have indicated that the turbulent kinetic energy (TKE) dissipation rate ε varies in the range of 10-9-10-5 W kg-1, corresponding to a diapycnal diffusivity $\kappa$ of 10-6-10-2 m2 s-1. However, measurements of turbulence are absent in the NTI. Using ship-based observations from late June to early July in the summer of 2006, this paper presents the first observations of turbulent mixing associated with the surface and internal tides in the NTI.

    The rest of this paper is organized as follows: In section 2, the data and methods are briefly introduced. In section 3, the ε and $\kappa$ in the NTI are statistically analyzed based on observations. Internal tides and the associated turbulent mixing are discussed in section 4. Section 5 is used to explain the enhancement of turbulence on the shelf of the NTI. The Ekman dynamics are also documented in this section. In section 6, the tidal modulation of water masses in the NTI are discussed. A summary and conclusion are given in section 7.

2. Data and methods
  • A cruise survey along two cross-shelf sections A and B in the NTI was conducted by R/V Dong Fang Hong 2 in the summer of 2006 (Fig. 1 and Table 1). The cruise started from station A01 on 28 June and ended at station B01 on 3 July. Combined conductivity-temperature-depth (CTD) and lowered acoustic Doppler current profilers (LADCP) measurements were carried out at 18 sites along transects A and B. Following the CTD/LADCP measurements, the Turbulence Ocean Microstructure Acquisition Profiler (TurboMAP-II, Alec Inc., Japan) was deployed to collect turbulence measurements. Stations A01-A05 and B01-B09 were on the shelf, while stations A06-A08 and B10 were on the slope, where the water depth is greater than 200 m. A total of 48 profiles of CTD and LADCP deployments at two cross sections were obtained. Stations A02 (anchored) and A06 (not anchored) were two 25-h-duration repeat stations. The time interval of each CTD/LADCP casting at A02 and A06 was 1.5 h and 2 h, respectively. At S06, the vessel was not anchored because the water depth exceeded 1000 m, but the vessel returned to the designated location after each deployment.

    A downward-looking 300-kHz ADCP was mounted on the CTD rosette package. The vertical bin size was set to 8 m and the sampling frequency was 1 Hz. To obtain accurate current data, the falling velocity of the instrument was controlled to be no more than 0.5 m s-1. The relative horizontal velocity was measured throughout the water column with an estimated uncertainty of 1 cm s-1, and the absolute velocity was derived from the relative velocity by a linear inverse method (Visbeck, 2002). The raw LADCP data were processed using an updated MATLAB package from Martin Visbeck (personal communication, 2008). The synchronized CTD time series, GPS position and bottom track velocity data were also used to calculate the pressure, vessel velocity and absolute bottom velocity.

    The free-falling TurboMAP-II measured small-scale velocity shear and temperature microstructure at a descending speed of 0.5 m s-1, as well as conductivity and pressure. The range of the measured change in velocity was from 0.001 to 0.5 m s-2, with a precision of 5%, while the precisions of the temperature and conductivity measurements were 0.01°C and 2× 104 m s-1 cm-1, respectively. The velocity microstructure shear was used to compute the ε, with a noise level of 10-10 W kg-1. The operating principle, calibration, signal handling, and limitations were described in (Wolk et al., 2002). The maximum measurement depth was set to 500 m. Five microstructure profiles were obtained at A02 (two profiles), A06 (two profiles) and A07 (one profile) in section A. In section B, nine TurboMAP-II profiles were obtained from B09 to B01 (Table 1). A more detailed description and technical report on these fine-structure profiles in the ECS can be found in the thesis of (Zhong, 2009).

  • 2.2.1. Harmonic analysis

    In the ECS, the principal tidal constituent is the semidiurnal M2 tide (Baines, 1982; Guo and Yanagi, 1998; Niwa and Hibiya, 2004; Garrett and Kunze, 2006). The diurnal K1 tide is also significant in this region (Zhu and Liu, 2012). To obtain the K1 and M2 tidal currents u K1, M2(z,t), harmonic analysis was applied to the observed velocity (u,v). The realistic depth at A06 exceeded 1000 m (Fig. 1b and Table 1). However, the maximum measurement depth in some profiles reached only 360 m (not anchored). Therefore, harmonic analysis was applied to the observed velocities only at depths above 360 m. Harmonic analysis does not account for incoherent internal tides (Kerry et al., 2014). The incoherent portion is negligible for a 25-h period.

    2.2.2. Estimation of dissipation rate

    With the assumption of isotropy, ε was computed by integrating the measured shear spectrum from TurboMAP-II: \begin{equation} \varepsilon=\dfrac{15}{2}\nu\overline{\left(\dfrac{\partial V}{\partial z}\right)^2}=\dfrac{15}{2}\nu\int_{k_1}^{k_2}\psi(k){\rm d}k , \ \ (1)\end{equation} where Ν is the coefficient of kinematic viscosity, ψ is the shear spectrum, and the overbar denotes the average depth. The small-scale velocity shear ∂ V/∂ z is estimated using Taylor's frozen assumption, \begin{equation} \dfrac{\partial V}{\partial z}=\dfrac{1}{W}\dfrac{\partial V}{\partial t} , \ \ (2)\end{equation} where V is the small-scale velocity and W is the sinking velocity of the instrument. Equation (2) was applicable when the sinking velocity W of TurboMAP-II exceeded 0.3 m s-1. On the right-hand side of Eq. (1), ψ is integrated over the wavenumber k. The wavenumber k1 in Eq. (1) was set to 1 cpm (circle per meter) and the upper limit k2 is the cutoff wavenumber that is not contaminated by high-frequency noises. For each profile, the Nasmyth spectrum (Nasmyth, 1970) was routinely used to evaluate the shapes of the measured dissipation spectra (Zhong, 2009).

    Based on ε, the diffusivity $\kappa$ can be calculated by \begin{equation} \kappa=\lambda\dfrac{\varepsilon}{N^2} . \ \ (3)\end{equation} When ε/(Ν N2)>100 (Ν is the molecular viscosity), Λ=2[ε/(Ν N2)]-1/2; otherwise, Λ=0.2 (Osborn, 1980; Shih et al., 2005). Equation (3) is applicable only when the buoyancy frequency N2>0. To rule out the effects of the vessel on turbulence, we focused only on the results below the depth of 10 m.

3. Microstructure observations from Turbo-MAP-II
  • Table 2 lists the mean, maximum and minimum ε and $\kappa$ values at all available fixed sites along transects A and B. The ε in the NTI varied from 10-9 to 10-5 W kg-1, which was comparable to previous observations in other parts of the ECS. The depth-averaged ε on the slope in section A was 1.0× 10-7 W kg-1, and the maximum ε was 3.7× 10-6 W kg-1. On the shelf, ε had a maximum of 7.9× 10-5 W kg-1, and its depth-average was 1.3× 10-6 W kg-1. The depth-mean $\kappa$ on the slope had an order of 10-4 m2 s-1, which was two orders smaller than that on the shelf.

    Figure 2 shows the distribution of density, ε and $\kappa$ along transect A. The depth-mean ε and $\kappa$ at A02 were 10-6 W kg-1 and 10-2 m2 s-1, respectively. The ε showed a significant bottom enhancement at A02, with a maximum value of 3.9× 10-5 W kg-1. Although there was strong stratification at A06 and A07, large ε and $\kappa$ values were observed in the pycnocline at A06, and the enhanced ε at A07 appeared at depths of 150 to 250 m. Along transect B, a surface mixed layer of ∼15 m was observed (Fig. 3). An enlarged ε often occurred in the upper 40 m from B01 to B07 (Fig. 3a). Below 40 m, there was also sporadic strong turbulence. At B01, the turbulence was enhanced in the bottom mixed layer (below 20 m). As a result, very large $\kappa$ (>10-2 m2 s-1) was observed therein (Fig. 3b). At B08 and B09, close to the continental slope, large ε and $\kappa$ were observed near the pycnocline.

    Figure 2.  The sectional density fields in section A incorporated with the ε [log10$\varepsilon$; (a); black curve] and vertical turbulent mixing [log10$\kappa$; (b); black curve] at A02, A06 and A07. The interval of the density contours is 0.5 kg m-3.

    Figure 3.  The ε [lg$\varepsilon$; (a)] and vertical turbulent mixing [lg$\kappa$; (b)] incorporated with the density (units: kg m-3; black contours) fields along transect B. The interval of the density contours is 1 kg m-3.

4. Continuous current and hydrological observations
  • Near the continental slope, enhanced stratified turbulence was observed along both transects A and B. In this section, we present the continuous observations from A06 to explore the possible mechanisms for enhanced stratified turbulence. As a comparison, another set of continuous observational records at a fixed site on the shelf (A02) is also analyzed.

  • Figures 4a and d show the profiles of time-mean velocities, $\bar{u}(h)$ and $\bar{v}(h)$, at A02 and A06, respectively. The TWC persists almost throughout the year with a transport of ∼ 1.5 Sv, providing a source- and sink-driven dynamic for the warm current in the South China Sea (Su and Wang, 1987; Su et al., 1994; Yang, 2007; Yang et al., 2008). At A02, the TWC was observed as flowing northeast at a maximum velocity of 0.4 m s-1 (Fig. 4a). The northward component $\bar{v}$ remained at 0.2 m s-1 throughout the water column, while the eastward component $\bar{u}$ was 0.3 m s-1 in the upper 30 m but decreased to 0.1 m s-1 from 30 to 75 m. At A06, the upslope currents had longer durations than the off-shelf currents, providing a net onshore transport (Fig. 4d). A significant on-shelf Kuroshio intrusion with a dominant westward velocity component (u=0.2 m s-1) was observed from 100 m to 250 m. In the upper layer (above 100 m), the velocity was nearly northward (v=0.3 m s-1). Beneath 250 m, the zonal velocity had a strong shear.

    Figure 4.  The time-mean zonal [$\bar{u}$(h); m s-1; red line] and meridional [$\bar{v}$(h); m s-1; blue line] currents at (a) A02 and (d) A06. The observed full (original), zonal [u(h,t); m s-1] and meridional [v(h,t); m s-1] currents at (b, c) A02 and (e, f) A06. The contour interval was set to 0.1 m s-1. The zero-contour of the velocity is highlighted by a thick black line. The times on the x-axis represent the deploying casts of CTD and LADCP. The red triangles at the bottom indicate the time of TurboMAP-II profiles.

    Figure 5.  The zonal [u bac(h,t); m s-1] and meridional [v bac(h,t); m s-1] baroclinic currents at (a, b) A02 and (c, d) A06. The contour interval was set to 0.05 m s-1 for A02 and 0.1 m s-1 for A06. The zero-contour of the velocity is highlighted by a thick black line. The bottom red triangles indicate the time of TurboMAP-II profiles. The isopycnal variations (σ; kg m-3) are incorporated by green contours with black labels.

    At both A02 and A06, the periodic tidal motions were visible. Semidiurnal barotropic tides were predominant at A02, with a maximum velocity of 1 m s-1 (Figs. 4b and c). The zonal velocity was larger than the meridional component. At A06, significant current shears were observed (Figs. 4e and f), suggesting the dominance of baroclinic motions on the slope.

    To distinguish internal tides from barotropic tides, the observed high-frequency velocity (u',v') was decomposed into barotropic and baroclinic components at A02 and A06: \begin{eqnarray} u'(h,t)&=&u-\bar{u}=u_{\rm bat}(t)+u_{\rm bac}(h,t) , \ \ (4)\\ v'(h,t)&=&v-\bar{v}=v_{\rm bat}(t)+v_{\rm bac}(h,t) , \ \ (5)\end{eqnarray} where u bat and v bat are the depth-mean (barotropic) velocity and u bac and v bac are the baroclinic velocity.

    The baroclinic currents at A02 were relatively weak, with a maximum velocity of 0.15 m s-1 (Figs. 5a and b). The isopycnals, especially the contour of 22.5 kg m-3, showed a semidiurnal fluctuation with the largest vertical displacement of ∼ 10 m, suggesting relatively weak internal tides. Large baroclinic velocities were observed at A06 (Figs. 5c and d), with a maximum velocity of 0.5 m s-1. The largest isopycnal displacement with a semidiurnal period was greater than 100 m. These observations suggested the occurrence of strong internal tides on the slope.

    Harmonic analysis was applied to u' and v' to obtain a time series of the tidal currents u K1+u M2 and v K1+v M2 (Fig. 6). At A02, the tidal currents showed an anticlockwise rotation with a dominant zonal component (u). Therefore, the tidal ellipses were stretched across the isobaths. At A06, two significant features were found. First, u K1+u M2 and v K1+v M2 had comparable magnitudes. The tidal ellipses on the slope (A06) were significantly different from those on the shelf (A02). Second, the tidal currents at A06 showed an evident baroclinic mode in the vertical profile, suggesting dominant internal tides.

    Figure 6.  The sum of the diurnal and semidiurnal currents (units: m s-1) by harmonic fit at (a, b) A02 and (c, d) A05. The contour interval is 0.1 m s-1. The zero-contour of the velocity is highlighted by a thick black line.

  • To understand how turbulence was generated, the vertical velocity shears (uz2+vz2) were calculated. Figure 7 shows shears in (u,v), $(\bar{u},\bar{v})$ (u',v') and (u K1+u M2,v K1+v M2) at A02 and A06. At A02, the high-frequency velocity shears dominated the full velocity shears from the surface to 45 m and below 70 m. From 50 m to 70 m, the shears in residual and high-frequency currents had comparable magnitudes, contributing to the full velocity shears. The high-frequency shears were primarily generated by currents K1 and M2. At A06, the shears in the upper 150 m (pycnocline) were approximately one order greater than those below 150 m. From the surface to 150 m, the total shears were dominated by the K1 and M2 current shears, while the total shears were dominated by the high-frequency shears below 150 m. The contribution of the residual current shears was negligible. In this paper, as the hourly observations in a day were not sufficient to decompose full velocity into all componential velocities with different frequencies, the high-frequency velocity shears include the shears associated with internal tides, the energy cascade through nonlinear wave-wave interaction (e.g., Xie et al., 2008), and/or some other physical mechanisms that may also play important roles in enhancing the energy dissipation.

    Figure 7.  The logarithms of time-averaged velocity shears (units: s-2) at (a) A02 and (b) A06. The observed full velocity shears are in red, the residual mean shears are in magenta, high-frequency velocity (full-residual) shears are in black, and K1+M2 shears are in blue. The vertical means of the above velocity shears are plotted in the same color at the bottom of the figures.

    To link turbulence to the shear instability associated with internal tides, the Richardson number Ri=N2/[(∂ u/∂ z)2+(∂ v/∂ z)2] was computed from the LADCP-based velocity shears and CTD-based buoyancy frequency. The ε and diffusivity showed similar vertical variations at both sites (Fig. 8). At S02-2, Ri was large ( Ri>1) between the depths of 20 and 30 m, corresponding to small ε and $\kappa$ values. The Ri decreased with increasing depth, and it was smaller than 0.25 near the bottom where the ε and $\kappa$ values were enhanced. This suggested that enhanced turbulence may be generated by shear instability. At A06, peaks of larger ε and $\kappa$ values were often accompanied with minimum Ri values. When Ri was less than 0.25, ε was elevated to 10-7 W kg-1 due to potential internal wave breaking. At these depths, $\kappa$ reached 10-3 m2 s-1. Because B08 was close to the continental slope, enhanced stratified turbulence was also observed in this area, as shown in Fig. 3.

    Figure 8.  The Richardson number Ri=N2/[(∂ u/∂ z)2+(∂ v/∂ z)2] (black), the logarithms of observed ε (W kg-1) (red) and turbulent mixing $\kappa$ (blue) in the vertical profile at (a) A02-2 and (b) A06-1. The Ri values larger than 5 were set to be 5 for visualization purposes. The TurboMAP-II was immediately deployed after the CTD/LADCP package returned. The observation time for A02-2 was 1123 UTC 29 June 2006, while it was 0837 UTC 30 June for A06-1. Note the different y-axis for the two figures.

5. Bottom boundary layer
  • Although the above observations indicated that the internal tides on the slope (A06) were much larger than those on the shelf (A02), the turbulence on the shelf was one to two orders of magnitude stronger than that on the slope. At A02, ε was enhanced below 40 m (Fig. 8a), where a thick bottom mixed layer (35 m) was observed (Fig. 9a) with dρ/dz less than 3× 10-3 kg m-4. Here, ρ represents the density of sea water. Similar observations also occurred at B01, where the thickness of the bottom mixed layers was ∼20 m (see Fig. 3a). Therefore, the enhanced turbulence at A02 and B01 may be attributable to the bottom boundary turbulence associated with the current shears caused by bottom friction. At A02, the shear instability associated with tidal currents and/or the residual currents (TWC) may explain the enhancement of turbulence (see Figs. 7a and 8a). However, weak turbulence was also observed at depths of 60-70 m at the bottom mixed layer at this site. Therefore, the boundary law-of-the-wall theory——namely, the inverse relationship between the bottom dissipation rate and distance from the seafloor (Trowbridge et al., 1999; Peters and Bokhorst, 2000)——cannot be directly used to explain the turbulence observed throughout the boundary layer. It is necessary for the rotary effect (Ekman) on the boundary layer, which may suppress turbulence (e.g., Perlin et al., 2007; Yoshikawa et al., 2010), to be taken into account.

    Figure 9.  (a). Vertical structure of density (σ; kg m-3) averaged over 17 profiles at A02. (b). Velocity magnitude (units: m s-1) of the mean velocity of the 17 profiles in 25 h at A02 (solid blue line with circles). (c) Current direction (units: °) of the mean velocity (solid red line with circles).

    The bottom boundary layer (BBL) on the shelf often consists of a thick Ekman layer over a thin logarithmic layer, which has been previously observed in the northern ECS (e.g., Yoshikawa et al., 2010). Beneath 45 m, the velocity magnitude associated with the northeastward TWC decreased gradually from 0.32 m s-1 in the interior to 0.18 m s-1 near the bottom (Fig. 9b). The rotation of the mean current spirals at A02 from the depth of 45 to the bottom was approximately 27° (Fig. 9c). This seems to go against classical Ekman theory, which predicts the rotation angle of the velocity veers anticlockwise by 45° in the Northern Hemisphere from the direction of the interior geostrophic velocity (Pedlosky, 1987). However, previous observations have indicated that the anticlockwise veering over the BBL is much less than 45°, typically reaching a maximum of only 20° (Perlin et al., 2007). The rotation angle of 27° in this paper is similar to the above observations.

    The mean observed diffusivity was used to calculate the viscosity based on the balance of the turbulence production and dissipation (Zhiyu LIU, personal communication, 2018). The turbulence shear production S=ρ Av[(∂ u/∂ z)2+(∂ v/∂ z)2] and buoyant production $B=\kappa\rho N^2$ are balanced by the TKE dissipation E=ρε, where Av is the viscosity. The balance is S-B=E. Using the calculation of $\kappa=\lambda\varepsilon/N^2$, where the mixing efficiency is 0.2, the buoyancy contribution was obtained as B=0.2ρε. Thus, the shear production S=ρ Av[(∂ u/∂ z)2+(∂ v/∂ z)2]=1.2ρε. That is, Av[(∂ u/∂ z)2+(∂ v/∂ z)2]=1.2ε. The vertical scale for the shears of ∂ u/∂ z and ∂ v/∂ z was approximately 0.9 m, which was determined from the spectrum analysis [see Eqs. (1) and (2)] based on standard technical procedure (Wolk et al., 2002). The observed ε and velocity shears can be used to obtain the viscosity Av. The mean viscosity was approximately 7× 10-3 m2 s-1. The observed vertical mean viscosity (7×10-3 m2 s-1) and interior velocity (0.3 m s-1) were used to calculate the classic Ekman solutions. From 50 m to 70 m, the predicted bottom Ekman currents agreed well with the observations in the velocity direction, but the magnitudes were slightly larger than observed (Fig. 10). At 75 m, the theoretical solution underestimated the real magnitude and direction. This underestimation may be attributable to the use of constant viscosity (Cushman-Roisin and Malačič, 1997; Perlin et al., 2005, Perlin et al., 2007). It has to be noted that the classic bottom Ekman theory was applied to this situation in which the vertical viscosity shows significant vertical and temporal variations. In theory, the bottom Ekman spiral can be modified slightly by the use of different vertical viscosity; however, the basic structure of bottom Ekman spiral does exist here. In this paper, we only depict and identify the Ekman dynamics and its dynamical connection with the local turbulent processes. The effects of different choices of viscosity on modifying the shapes of Ekman spirals are beyond the scope of this study and these discussions can be found in (Huang, 2010). On the other hand, such accumulative observations of Ekman spirals will help obtain the vertical viscosity inversely (Yoshikawa et al., 2010).

    Figure 10.  Observed velocity vectors (black arrows) and spirals in the BBL at A02 versus the analytical Ekman solutions (red curve with circles) using the observed magnitude of the TWC and bottom turbulent viscosity (Table 2). The numbers near the vectors and curves are the depths of the velocity. The dashed purple lines and numbers represent the velocity magnitudes. The MATLAB program of bottom Ekman solution was adapted from J. F Price's MIT-WHOI joint program class in 2009.

    Below 63 m, the dissipation rate at A02 showed an increase, agreeing with the law-of-the-wall theory for the estimation of the bottom dissipation rate ε b. Here, we also used this theory to estimate ε b (Trowbridge et al., 1999; Peters and Bokhorst, 2000). In this theory, the dissipation rate in the constant stress layer is related to the bottom friction velocity u*: \begin{eqnarray} \varepsilon_{\rm b}&=&u_*^3(\kappa_{\rm vk}z)^{-1} , \ \ (6)\\ u_*&=&\sqrt{\dfrac{\tau_{\rm b}}{\rho}} , \ \ (7) \end{eqnarray} where z is set to 4 m, $\kappa_{vk}$ (=0.4) is the von Kármán's constant, and τ b is the magnitude of the bottom shear stress: \begin{equation} \tau_{\rm b}=\rho c_{\rm d}|u_{\rm b}|u_{\rm b} , \end{equation} where u b is the LADCP bottom track velocity and c d (=0.002) is the drag coefficient.

    Figure 11 shows the near-bottom velocity and ε estimated by Eq. (6) at A02. The bottom velocities ranged from 0.1 m s-1 to 0.5 m s-1, with a semidiurnal period. The bottom dissipation rate ε b varied from 10-7 W kg-1 at flat tide to 10-5 W kg-1 at flood or ebb tide. During the TurboMAP-II deployments, the bottom currents were larger and ε b=10-5 W kg-1 (Fig. 11b), which was consistent with the direct microstructure observations at 73 m (Fig. 8a).

    Figure 11.  (a) The near-bottom currents from the bottom track of LADCP (at 75 m) during 28-29 June. (b) The estimated near-bottom ε [ log10ε b; units: W kg-1)] based on Eqs. (6) and (7). The bottom triangles indicate the times of the TurboMAP-II profiles.

6. Water mass properties associated with tides in the NTI: a discussion
  • Three water mass systems were observed during both transects A and B (Fig. 12): denser Kuroshio bottom water (KBW) characterized by low temperature (T) and high salinity (S), warm and saline Kuroshio surface water and subsurface waters (KSW), and shelf water (SW) with low S. The KBW was stable, with the potential density greater than 25 kg m-3. The density of KSW ranged from 21.5 kg m-3 to 25.0 kg m-3. The isohaline of 34.4 psu was used to separate the KSW and SW.

    The T-S property at both A06 and A02 had a tidal period modulation, indicating the role of the rotating surface tides (Fig. 12a). At flood tide, A02 featured high T and S, with an opposite situation at ebb tide (Figs. 13a and c). However, the effect of tidal advections on the T-S diagram was weak above 10 m and below 35 m. During the upslope transport (Figs. 13b and c), the T-S relationship at A06 was similar to that of KSW, while it approached the T-S characteristics of the SW during the downslope transport. However the effect of tidal advection on the T-S characteristics at A06 was confined to the upper 120 m (Fig. 13c).

    Figure 12.  The T-S diagrams for (a) cross-section A and (b) cross-section B, with the colors labeled on the top of the curves. The water masses of KBW, KSW and SW are indicated by dark blue, purple and dark red frames, respectively.

    Figure 13.  Comparisons of the T-S diagram at A02 and A06 with different directions of tidal currents. Red (black) curves and arrows in (a) and (c) represent the T-S properties and the offshore (onshore) direction currents at A02. Magenta (blue) curves and arrows in (b) and (c) represent the T-S property and the offshore (onshore) direction currents at A06. Panel (b) is the same as (a) but for A06. Orange circles and labels in (a) and (b) represent the velocity magnitude. The observation times of these profiles are labeled in (c).

7. Summary and conclusion
  • Using a set of hydrological and microstructure records collected in the summer of 2006, we investigated the tides and mixing in the NTI. The results showed that barotropic tides were dominant on the shelf but internal tides were significant on the slope, particularly in the upper layer from the surface to 150 m. On the shelf, strong turbulence with ε=10-5 W kg-1 was observed in a thick bottom boundary. The current shears caused by the bottom friction in the shallow area were essential for enhanced turbulence. A bottom Ekman spiral was also observed in the boundary layer, agreeing well with the analytical bottom Ekman solution. On the slope, the enhanced turbulence was primarily caused by shear instability associated with internal tides. The mixing rate $\kappa$ on the shelf (10-2 m2 s-1) was two orders of magnitude larger than that on the slope (10-4 m2 s-1).

Reference

Catalog

    /

    DownLoad:  Full-Size Img  PowerPoint
    Return
    Return