Advanced Search
Article Contents

The Impact of Deformation on Vortex Development in a Baroclinic Moist Atmosphere


doi: 10.1007/s00376-015-5082-y

  • A mathematical relation between deformation and vertical vorticity tendency is built by introducing the frontogenesis function and the complete vertical vorticity equation, which is derived by virtue of moist potential vorticity. From the mathematical relation, it is shown that properly configured atmospheric conditions can make deformation exert a positive contribution to vortex development at rates comparable to other favorable factors. The effect of deformation on vortex development is not only related to the deformation itself, but also depends on the current thermodynamic and dynamic structures of the atmosphere, such as the convective stability, moist baroclinicity and vertical wind shear (or horizontal vorticity). A diagnostic study of a heavy-rainfall case that occurred during 20-22 July 2012 shows that deformation has the most remarkable effect on the increase in vertical vorticity during the rapid development stage of the low vortex during its whole life cycle. This feature is mainly due to the existence of an approximate neutral layer (about 700 hPa) in the atmosphere where the convective stability tends to be zero. The neutral layer makes the effect of deformation on the vertical vorticity increase significantly during the vortex development stage, and thus drives the vertical vorticity to increase.
  • 加载中
  • Bishop C. H., 1996a: Domain-independent attribution. Part I: Reconstructing the wind from estimates of vorticity and divergence using free space green's functions. J. Atmos. Sci., 53, 241- 252.2c7de67aee697eb9351c088b24aec901http%3A%2F%2Fconnection.ebscohost.com%2Fc%2Farticles%2F9602270524%2Fdomain-independent-attribution-part-i-reconstructing-wind-from-estimates-vorticity-andhttp://connection.ebscohost.com/c/articles/9602270524/domain-independent-attribution-part-i-reconstructing-wind-from-estimates-vorticity-andPresents mathematical models for reconstructing wind dynamics from estimates of vorticity and divergence using free space Green's function. Reconstruction technique on discrete data; Accuracy of standard finite-difference estimates of vorticity/divergence; Attribution and wind field representation on Cartesian domains; Application to limited region of the sphere.
    Bishop C. H., 1996b: Domain-Independent attribution. Part II: Its value in the verification of dynamical theories of frontal waves and frontogenesis. J. Atmos. Sci., 53, 253- 262.10.1175/1520-0469(1996)053<0253:DIAPII>2.0.CO;24ce333d4-8238-4c0e-af9e-7524f06a47b217044750f95881eec71c1db228ae93b5http%3A%2F%2Fwww.researchgate.net%2Fpublication%2F252447633_Domain-Independent_Attribution._Part_II_Its_Value_in_the_Verification_of_Dynamical_Theories_of_Frontal_Waves_and_Frontogenesishttp://www.researchgate.net/publication/252447633_Domain-Independent_Attribution._Part_II_Its_Value_in_the_Verification_of_Dynamical_Theories_of_Frontal_Waves_and_FrontogenesisAbstract Theories of frontogenesis and frontal waves describe development in terms of the interaction of a basic state or environmental flow with a frontal flow. The basic-state flow may comprise a large-scale confluent iffluent deformation field and/or an alongfront temperature gradient. The frontal flow is seen as evolving as a result of its interaction with the environmental flow. Such theories make specific predictions about the effect of the basic-state flow on the frontal flow. To test these predictions, counterparts of the basic-state flows and frontal flows used in theoretical models must be extracted from atmospheric data. Here the concept of attribution is used to identify such counterparts. In the present context, attribution refers to the process whereby a part of the wind field is attributed to a part of the vorticity or divergence field. It is mathematically equivalent to the process by which a part of a field of electric potential is associated with an element of total charge density in electrostatics. The counterpart of the frontal flow used in idealized models is identified as that part of the flow attributable to the vorticity and divergence anomalies within the frontal region. The counterpart of the basic-state flow is identified as that part of the flow attributable to vorticity and divergence anomalies outside the frontal region. Applications of the partitioning method are illustrated by diagnosing the flow associated with a North Atlantic front. The way in which the partitioning method may be used to test some theories concerning the effect of large-scale deformation on frontal wave formation is described. The partitioning method's ability to distinguish frontogenesis due to environmental flow from that due to frontal flow is also discussed. The analyzed front is found to lie at an angle to the dilatation axis of the environmental flow. It is argued that this feature must be common to all nonrotating finite length fronts.
    Bishop C. H., A. J. Thorpe, 1994a: Frontal wave stability during moist deformation frontogenesis. Part I: Linear wave dynamics. J. Atmos. Sci., 51( 6), 852- 873.10.1175/1520-0469(1994)0512.0.CO;285c6f5bb-1341-43eb-bd59-54f788396889c26d8e95f2f874d77173bfefa7ed8df8http%3A%2F%2Fwww.researchgate.net%2Fpublication%2F249609557_Frontal_Wave_Stability_during_Moist_Deformation_Frontogenesis._Part_I_Linear_Wave_Dynamicsrefpaperuri:(12f4cb980b34e15ffe642bf52248033e)http://www.researchgate.net/publication/249609557_Frontal_Wave_Stability_during_Moist_Deformation_Frontogenesis._Part_I_Linear_Wave_DynamicsAbstract It has been shown that lower tropospheric potential vorticity zones formed during moist deformation frontogenesis will support growing waves if at some time the frontogenesis ceases. In this paper, the ways in which these waves are affected by the frontogenetic process are identified. Observations show that fronts in the eastern Atlantic commonly feature saturated ascent regions characterized by zero moist potential vorticity. Furthermore, in many cases the horizontal temperature gradient in the lowest one to two kilometers of the atmosphere is rather weak. These features are incorporated in an analytical archetype. The dynamical implications of saturated ascent in conditions of zero moist potential vorticity are represented in the model by assuming that adiabatic temperature changes are precisely balanced by diabatic tendencies. The observed small temperature gradient at low levels is represented in the model by taking it to be zero in the lowest two kilometers. Consequently, the forcing of the low-level moist ageostrophic vortex stretching that strengthens the low-level potential vorticity anomaly is confined to middle and upper levels. A semianalytical initial value solution for the linear development of waves on the evolving low-level potential vorticity anomaly is obtained. The waves approximately satisfy the inviscid primitive equations whenever the divergent part of the perturbation is negligible relative to the rotational part. The range of nonmodal wave developments supported by the front is summarized using RT phase diagrams. This analysis shows that the most dramatic effects of frontogenesis on frontal wave growth are due to (a) the increase in time of the potential vorticity and hence potential instability of the flow and (b) the increase in time of the alongfront wavelength relative to the width of the strip. An optimally growing streamfunction wave is described. Finally, a diagnostic technique suitable for identifying small amplitude frontal waves in observational data is described.
    Bishop C. H., A. J. Thorpe, 1994b: Frontal wave stability during moist deformation frontogenesis. Part II: The suppression of nonlinear wave development. J. Atmos. Sci., 51( 6), 874- 888.c76e606cc56d3f1a0c3a32245d176832http%3A%2F%2Fconnection.ebscohost.com%2Fc%2Farticles%2F9501264345%2Ffrontal-wave-stability-during-moist-deformation-frontogenesis-part-ii-suppression-ofhttp://connection.ebscohost.com/c/articles/9501264345/frontal-wave-stability-during-moist-deformation-frontogenesis-part-ii-suppression-ofAssesses the role of horizontal deformation and the associated frontogenetic ageostrophic circulation in suppressing the development of nonlinear waves. Wave slope and its maximal amplification; Parameter space of maximal amplification; Description and interpretation of the effect of strain on growth; Scale selection.
    Bluestein H. B., 1993: Synoptic Dynamic Meteorology in Midlatitudes. Oxford University Press,594 pp.997641b78b6bc966943a20f660d37e09http%3A%2F%2Fci.nii.ac.jp%2Fncid%2FBA17138163http://ci.nii.ac.jp/ncid/BA17138163This new, comprehensive textbook for upper-division undergraduate and graduate students of meteorology presents for the first time information that is now considered essential in modern weather forecasting. Based on a successful series of courses taught by the author at the University of Oklahoma, the text carefully examines the foundations of synoptic meteorology, from the analysis of scalar fields to atmospheric kinematics, dynamics, and thermodynamics.
    Cai M., 1992: A physical interpretation for the stability property of a localized disturbance in a deformation flow. J. Atmos. Sci., 49( 23), 2177- 2182.10.1175/1520-0469(1992)049<2177:APIFTS>2.0.CO;281a0fb4335ad594fc052d11ba3968786http%3A%2F%2Fwww.researchgate.net%2Fpublication%2F249609372_A_physical_interpretation_for_the_stability_property_of_a_localized_disturbance_in_a_deformation_flowhttp://www.researchgate.net/publication/249609372_A_physical_interpretation_for_the_stability_property_of_a_localized_disturbance_in_a_deformation_flowABSTRACT
    Cai M., M. Mak, 1990: On the basic dynamics of regional cyclogenesis. J. Atmos. Sci., 47( 12), 1417- 1442.10.1175/1520-0469(1990)0472.0.CO;2d84faed4-addb-44fc-9533-64310c9ff433afb7d47864e0dfc69338a28528a6ca79http%3A%2F%2Fwww.researchgate.net%2Fpublication%2F234530620_On_the_Basic_Dynamics_of_Regional_Cyclogenesisrefpaperuri:(80b600b5735a23a080817e7b41b18624)http://www.researchgate.net/publication/234530620_On_the_Basic_Dynamics_of_Regional_CyclogenesisAbstract This paper investigates the dynamics of regional cyclogenesis from the perspective of local instability of a zonally inhomogeneous baroclinic jet streak in a two-layer quasi-geostrophic beta-plane channel model. When such a representative jet streak is embedded in a background uniform vertical shear U T , there are both local and global unstable normal modes. In the absence of such a background shear ( U T = 0), only the local modes are unstable. The shorter the jet is, the fewer local modes would there be. A local mode consists of a group of dominant waves that jointly give rise to a maximum local energy downstream of the jet core. Its existence is independent of the cyclical boundary condition. The growth rate of a local mode diminishes rapidly when the constant part of the basic zonal wind U 0 is increased. A global mode, on the other hand, largely consists of a single wave and its growth rate is much less sensitive to U 0 . These properties are qualitatively similar to those in the WKB solution. The structural characteristics of these modes are identifiable with those of three classes of unstable modes of an observed atmospheric flow reported in Frederiksen and Bell. Our nonmodal analysis shows that a localized disturbance naturally emerges from a zonally unbiased initial state in a relatively short time. The excitation of a local mode within a few days from an initially isolated disturbance also depends strongly upon its initial position relative to the jet core. The two processes that locally generate the perturbation energy depend upon the structural properties of the disturbance relative to the basic thermal and deformation fields. The two processes that redistribute the perturbation energy are the advection of energy by the basic flow and the convergence of energy flux associated with the ageostrophic component of the perturbation. These four processes are comparably important and greatly counteract one another resulting in a net intensification of a disturbance centered downstream of the jet core. The feedback effects of the most unstable mode on the basic flow resemble the observed geopotential tendencies induced by the transient eddies. The feedback results of this analysis differ noticeably from the WKB counterparts.
    Cui X. P., G. X. Wu, and S. T. Gao, 2002: Numerical simulation and isentropic analysis of frontal cyclones over the western Atlantic Ocean. Acta Meteorologica Sinica, 60( 4), 385- 399. (in Chinese)10.1002/mop.105028c3370f6-396e-4603-99af-134c16f8379f55842002486ede3cfd43831887b2a77cfb7104a53ahttp%3A%2F%2Fen.cnki.com.cn%2FArticle_en%2FCJFDTOTAL-QXXB200204000.htmhttp://en.cnki.com.cn/Article_en/CJFDTOTAL-QXXB200204000.htmFrontal cyclogenesis is a widespread weather phenomenon. In spite of the marked improvements in the study of this kind of phenomenon, many problems ramain. Our understanding and prediction of this type of mesoscale vortex is still hampered by (1)the lack of high-resolution data and (2)the absence of theoretical models. This kind of understanding and forecast is still among the most serious challenges to atmospheric scientists. Here by using PSU/NCAR MM5 mesoscale model, a 60-h simulation is performed to reproduce a frontal cyclogenesis over the Western Atlantic Ocean during March 13-15, 1992. Beginning with Slantwise Vorticity Development(SVD), the genesis ,development and propagation are studied by using high-resolution model output in the context of slantwise isentropic surface. And inspiring results are received. The model reproduces well the genesis, track and intensity of the cyclones, their associated thermal structure as well as their surface circulation. The major cyclone(M) deepens 45hPa in the all 60 h and 12hPa in 6 h from 36 to 42(model time) and 27hPa in 24 h from 36 to 60(model time). Profile and isentropic analysis tell us that the cyclogenesis is in very close relation with slantwise isentropic surface; The cyclones always superpose with the core of neutral convection stability with nearly vertical isentropic surface, which coincides with what SVD says.Beginning with Slantwise Vorticity Development(SVD),the development and propagation of the oceanic frontal cyclone are studied by using high-resolution model output in the context of slantwise isentropic surface.The results show that the frontal cyclone deepens rapidly by interaction with large-scale environment after occurring over ocean with weak hydrostatic stability; SVD theory can well translate the development and propagation, which is closely related with slantwise isentropic surface. The downstream slantwise upsliding movement and declination of isentropic surface make vorticity develop(USVD) under favorable conditions ( C D0, here C D is SVD index),and result in the moving and development of cyclone.
    Charney J. G., 1947: The dynamics of long waves in a baroclinic westerly current. J. Meteor., 4, 136- 162.10.1175/1520-0469(1947)0042.0.CO;2ec38cceb1cb5771f96e1fabd0a960fadhttp%3A%2F%2Fwww.ams.org%2Fmathscinet-getitem%3Fmr%3D22136http://www.ams.org/mathscinet-getitem?mr=22136Abstract Previous studies of the long-wave perturbations of the free atmosphere have been based on mathematical models which either fail to take properly into account the continuous vertical shear in the zonal current or else neglect the variations of the vertical component of the earth's angular velocity. The present treatment attempts to supply both these elements and thereby to lead to a solution more nearly in accord with the observed behavior of the atmosphere. By eliminating from consideration at the outset the meteorologically unimportant acoustic and shearing-gravitational oscillations, the perturbation equations are reduced to a system whose solution is readily obtained. Exact stability criteria are deduced, and it is shown that the instability increases with shear, lapse rate, and latitude, and decreases with wave length. Application of the criteria to the seasonal averages of zonal wind suggests that the westerlies of middle latitudes are a seat of constant dynamic instability. The unstable waves are similar in many respects to the observed perturbations: The speed of propagation is generally toward the east and is approximately equal to the speed of the surface zonal current. The waves exhibit thermal asymmetry and a westward tilt of the wave pattern with height. In the lower troposphere the maximum positive vertical velocities occur between the trough and the nodal line to the east in the pressure field. The distribution of the horizontal mass divergence is calculated, and it is shown that the notion of a fixed level of nondivergence must be replaced by that of a sloping surface of nondivergence. The Rossby formula for the speed of propagation of the barotropic wave is generalized to a baroclinic atmosphere. It is shown that the barotropic formula holds if the constant value used for the zonal wind is that observed in the neighborhood of 600 mb.
    Davies H. C., J. C. Müller, 1988: Detailed description of deformation-induced semi-geostrophic frontogenesis. Quart. J. Roy. Meteor. Soc., 114, 1201- 1219.10.1002/qj.49711448303386ed13fdb8409aba9d21b24db230c7ehttp%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1002%2Fqj.49711448303%2Fcitedbyhttp://onlinelibrary.wiley.com/doi/10.1002/qj.49711448303/citedbyA variant of the classical deformation-induced surface frontogenesis problem is studied in the semigeostrophic limit. The particular flow system examined is the response of a baroclinic, semi-infinite, stratified, uniform potential vorticity atmosphere to an imposed deformation flow field.
    Deng Q. H., 1986: The deformation field in the planetary boundary layer and heavy rainfall. Journal of Academy of Meteorological Science, 1, 165- 174. (in Chinese)a2f179a9359d49c55970efe5410ab864http%3A%2F%2Fen.cnki.com.cn%2FArticle_en%2FCJFDTotal-YYQX198602006.htmhttp://en.cnki.com.cn/Article_en/CJFDTotal-YYQX198602006.htmIn this paper,a statistical analysis was made on heavy rainfalls occurring in HubeiProvince and the deformation field in the planetary boundary layer during Meiyu period of1979—1981.It is found that heavy rainstorms are closely related to deformation field inthe planetary boundary layer.Analyzing the deformation field in the planetary boundarylayer gives a criterion for the movement and development of low pressure systems aswell as an indication for the forecasting of heavy rain area in 12—24hours.
    Eady E. T., 1949: Long waves and cyclone waves. Tellus, 1, 33- 52.10.3402/tellusa.v1i3.8507db89b19bc99767c1a89ac7525c7b4d31http%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1111%2Fj.2153-3490.1949.tb01265.x%2Fabstracthttp://onlinelibrary.wiley.com/doi/10.1111/j.2153-3490.1949.tb01265.x/abstractAbstract By obtaining complete solutions, satisfying all the relevant simultaneous differential equations and boundary conditions, representing small disturbances of simple states of steady baroclinic large-scale atmospheric motion it is shown that these simple states of motion are almost invariably unstable. An arbitrary disturbance (corresponding to some inhomogeneity of an actual system) may be regarded as analysed into &ldquo;components&rdquo; of a certain simple type, some of which grow exponentially with time. In all the cases examined there exists one particular component which grows faster than any other. It is shown how, by a process analogous to &ldquo;natural selection&rdquo;, this component becomes dominant in that almost any disturbance tends eventually to a definite size, structure and growth-rate (and to a characteristic life-history after the disturbance has ceased to be &ldquo;small&rdquo;), which depends only on the broad characteristics of the initial (unperturbed) system. The characteristic disturbances (forms of breakdown) of certain types of initial system (approximating to those observed in practice) are identified as the ideal forms of the observed cyclonc waves and long waves of middle and high latitudes. The implications regarding the ultimate limitations of weather forecasting are discussed.
    Elhmaidi D., A. Provenzale, T. Lili, and A. Babiano, 2004: Stability of two-dimensional vorticity filaments. Physics Letters A, 333, 85- 90.10.1016/j.physleta.2004.10.03319954e44027e5efc59b64270747d9d41http%3A%2F%2Fwww.sciencedirect.com%2Fscience%2Farticle%2Fpii%2FS0375960104014768http://www.sciencedirect.com/science/article/pii/S0375960104014768We discuss the results of a numerical study on the stability of two-dimensional vorticity filaments around a circular vortex. We illustrate how the stability of the filaments depends on the balance between the strain associated with the far field of the vortex and the local vorticity of the filament, and we discuss an empirical criterion for filament stability.
    Farrell B. F., 1989: Transient development in confluent and diffluent flow. J. Atmos. Sci., 46, 3279- 3288.10.1175/1520-0469(1989)046<3279:TDICAD>2.0.CO;26165bcb60a3c4d920e027f24f4bb0a5ahttp%3A%2F%2Fwww.ams.org%2Fmathscinet-getitem%3Fmr%3D1021134http://www.ams.org/mathscinet-getitem?mr=1021134Abstract Explaining the growth of disturbances superimposed on mean flows is a central problem in meteorology. Most widely studied models of the development process involve perturbations to shear flows with shear restricted to the meridional direction. Recently the importance of zonal variation of the mean flow was recognized and the study of shear flows extended to include zonal variation in shear. These studies found that the eigenfunctions associated with unstable modes in the simple shear problem are highly sensitive to zonal variation of the mean flow. However, there also exists another mechanism for development in a zonally inhomogeneous flow field: transient growth not associated with exponential instability. Properly configured perturbations exhibit transient growth in deformation fields associated with regions of confluence and diffluence at rates comparable to development in shear flow. In this work analytic solution of the linear initial value problem for the barotropic vorticity equation in deformation flow is used to construct local perturbations that undergo rapid transient development. Implications for cyclogenesis and block formation are discussed.
    Gao S. T., S. Yang, M. Xue, and C. M. Cui, 2008: Total deformation and its role in heavy precipitation events associated with deformation-dominant flow patterns. Adv. Atmos. Sci.,25(1), 11-28, doi: 10.1007/s00376-008-0011-y.10.1007/s00376-008-0011-ye383b57ebf09e15a790463659720f7cbhttp%3A%2F%2Fd.wanfangdata.com.cn%2FPeriodical_dqkxjz-e200801002.aspxhttp://d.wanfangdata.com.cn/Periodical_dqkxjz-e200801002.aspx
    Holton J. R., 2004: Circulation and vorticity. An Introduction to Dynamic Meteorology, Elsevier Academic Press, 84- 114.10.1088/0034-4885/5/1/3094a9a7c5124abd58d53e17ed1661afa88http%3A%2F%2Fwww.sciencedirect.com%2Fscience%2Farticle%2Fpii%2FB9780750663151500166http://www.sciencedirect.com/science/article/pii/B9780750663151500166Several charts are presented that show meteorology data and figures, including annual climatological data for U.S. cities in 2008, record temperature by U.S. state, and hurricane and tornado classifications.
    Hoskins B. J., F. P. Bretherton, 1972: Atmospheric frontogenesis models: Mathematical formulation and solution. J. Atmos. Sci., 29, 11- 37.10.1175/1520-0469(1972)0292.0.CO;2583547bc04a775d3c865a1ee9f6dada4http%3A%2F%2Fwww.researchgate.net%2Fpublication%2F243775183_Atmospheric_Frontogenesis_Models_Mathematical_Formulation_and_Solutionhttp://www.researchgate.net/publication/243775183_Atmospheric_Frontogenesis_Models_Mathematical_Formulation_and_SolutionAbstract The approximation of geostrophic balance across a front is studied. Making this approximation, an analytic approach is made to a frontogenesis model based on the classic horizontal deformation field. Kelvin's circulation theorem suggests the introduction of a new independent variable in the cross-front direction. The problem is solved exactly for a Boussinesq, uniform potential vorticity fluid. Non-Boussinesq, non-uniform potential vorticity, latent heat, and surface friction effects are all studied. Using a two-region fluid we model the effects of confluence near the tropopause. A similar approach is made to the appearance of fronts in the finite-amplitude development of the simplest Eady wave; this is also solved analytically. Based on the surface fronts produced by these models, we give a general model of a strong surface front. There is a tendency to form discontinuities in a finite time.
    Jiang Y. Q., 2011: Study on the dynamic mechanism of formation of mesoscale weather systems triggered by wind perturbations. Ph.D dissertation, Nanjing University, 156 pp.
    Jiang Y. Q., Y. Wang, Z. G. Zhou, M. Lü, and J. Luo, 2011: Interaction index of vortex and deformation field. Journal of PLA University of Science and Technology (Natural Science Edition), 12( 6), 685- 689. (in Chinese)87945211899db4de65977c82eefec68ahttp%3A%2F%2Fen.cnki.com.cn%2FArticle_en%2FCJFDTOTAL-JFJL201106026.htmhttp://en.cnki.com.cn/Article_en/CJFDTOTAL-JFJL201106026.htmTo analyze the short-term change of the intensity and the moving track and speed of the vortex in the deformation field,an interaction index of vortex and deformation field(VDI) was presented.The background pattern of a heavy rainfall event caused by a tropical depression(TD) during 5-6 August 2001 was analyzed with a digital filter compositing method based on Lanczos window function.It is shown that a col field forms as the strong Western Pacific subtropical high is splitting over eastern China.The TD intensifies as it is moving into the col field.Because the higher VDI which is favor of the development of TD represents the stronger interaction of the TD and the col field,the TD may move toward the high VDI region,so the VDI center can be taken as an indication of the moving direction of the TD.Furthermore,the distance between the TD center and the VDI center is also an indication of the moving speed of the TD.The bigger the size of distance,the quicker the TD moves.
    Keyser D., R. A. Anthes, 1982: The influence of planetary boundary layer physics on frontal structure in the Hoskins-Bretherton horizontal shear model. J. Atmos. Sci., 39, 1783- 1802.10.1175/1520-0469(1982)039<1783:TIOPBL>2.0.CO;2106a5d3ff77d233047e9c52fa5f2f147http%3A%2F%2Fonlinelibrary.wiley.com%2Fresolve%2Freference%2FADS%3Fid%3D1982JAtS...39.1783Khttp://onlinelibrary.wiley.com/resolve/reference/ADS?id=1982JAtS...39.1783KA series of numerical experiments with the Hoskins-Bretherton horizontal shear model of frontogenesis in an, amplifying, two-dimensional baroclinic wave is performed. The analytic solutions from the Boussinesq, semi-geostrophic model provide initial conditions for numerical integrations with a two-dimensional, dry version of the fully compressible, hydrostatic primitive equation (PE) model of Anthes and Warner with 40 km horizontal resolution. The PE model is integrated 1) without planetary boundary layer (PBL) physics; 2) with a one-layer bulk-drag scheme; and 3) with a high-vertical-resolution PBL model. The lower boundary is thermally insulated in order to isolate the effect of the internal mixing of heat in the PBL.The simulation with the high-resolution PBL physics resolves several realistic features including 1) a narrow updraft at the top of the PBL above the sea-level pressure trough at the warm edge of the frontal zone; 2) a stable layer capping the PBL to the rear of the frontal zone; and 3) slightly unstable or neutral lapse rates in the PBL behind the front and stable lapse rates in the PBL ahead of the front. A diagnostic analysis of the frontogenesis indicates that the fine structure resulting from adding PBL physics can be attributed to the frictionally driven, ageostrophic inflow in the PBL toward the surface pressure trough in which the frontal zone is located. A finding of particular interest is that the stability patterns in the PBL on either side of the front evolve independently of sensible heating at the surface.
    Keyser D., M. J. Reeder, and R. J. Reed, 1988: A Generalization of Petterssen's Frontogenesis Function and Its Relation to the Forcing of Vertical Motion. Mon. Wea. Rev., 116, 762- 781.10.1175/1520-0493(1988)1162.0.CO;270644bb348f1c6560b1ea5d2e8f298achttp%3A%2F%2Fwww.researchgate.net%2Fpublication%2F23873940_A_generalization_of_Petterssen%27s_frontogenesis_function_and_its_relation_to_the_forcing_of_vertical_motionhttp://www.researchgate.net/publication/23873940_A_generalization_of_Petterssen's_frontogenesis_function_and_its_relation_to_the_forcing_of_vertical_motionAbstract Petterssen' frontogenesis equation relates the Lagrangian rate of change of the magnitude of the horizontal potential temperature gradient, referred to as the frontogenesis function, to invariant kinematic properties of the horizontal velocity field. It is not uncommon in synoptic practice to infer the presence of vertical circulations in frontal regions from the spatial distribution of the scalar frontogenesis function. On the other hand, Hoskins and collaborators have introduced a form of the quasi-geostrophic omega equation in which the dynamical and forcing is proportional to the horizontal divergence of the so-called Q vector. The Q vector is defined as the Lagrangian rate of change following the geostrophic flow of the vector horizontal potential temperature gradient. The Q -vector formalism motivates us to generalize the Petterssen frontogenesis function to apply to the vector horizontal potential temperature gradient. This generalization, referred to as the vector frontogenesis function, consists of introducing an expression for the Lagrangian rate of change of direction of the horizontal potential temperature gradient. In order to investigate quantitatively the relative importance of the magnitude and direction contributions to the vector frontogenesis function, we consider three analytical examples. These examples describe the evolution of a potential temperature field represented initially by a linear band of isentropes situated within specified horizontal wind fields that are nondivergent and steady state. The wind fields respectively are a hyperbolic streamline pattern characterized by pure deformation, a meridional wind field varying only in the zonal direction, and an axisymmetric vortex. In each of these examples, it is found that the Lagrangian rates of change of the magnitude and direction of the potential temperature gradient are comparable. In order to explore the dynamical implications of this finding, we separate the Q -vector forcing into contributions consisting of the magnitude and direction components of the vector frontogenesis function. The outcome of this partitioning suggests a possible dynamical basis for isolating vertical circulations associated with frontal zones in three-dimensional baroclinic disturbances: the frontal circulation is related to the magnitude component of the Q vector, whereas the background circulation (that associated with the baroclinic disturbance) is related to the direction component. Consequently, the proposed partitioning of the Q vector appears to lend dynamical support to adopting the scalar frontogenesis function as a qualitative indicator of frontal circulations, provided that these circulations are understood to constitute only a component of the total vertical motion field.
    Kuo H. L., 1949: Dynamics instability of two-dimensional non-divergent flow in a barotropic atmosphere. J. Meteor., 6, 105- 122.c7db1477-b194-468f-b378-d9ba1b7395b1cd0ea3d06082e05aee3e721b7db86b90http%3A%2F%2Fwww.researchgate.net%2Fpublication%2F265981594_Barotropic_instability_of_two-dimensional_non-divergent_flow_in_the_atmospherehttp://www.researchgate.net/publication/265981594_Barotropic_instability_of_two-dimensional_non-divergent_flow_in_the_atmosphere
    Li Y., L. S. Chen, and X. T. Lei, 2005: Moisture potential vorticity analysis on the extratropical transition processes of Winnie (1997) and Bilis (2000). Journal of Tropical Meteorology, 21( 2), 142- 152. (in Chinese)10.1360/biodiv.05002860d15312-86b6-4482-8d24-e6d92eabb6bace81c95b963bd16252998557052b762chttp%3A%2F%2Fen.cnki.com.cn%2FArticle_en%2FCJFDTotal-RDQX200502004.htmrefpaperuri:(9592b3170ea89d982756150a69abc8ed)http://en.cnki.com.cn/Article_en/CJFDTotal-RDQX200502004.htmTyphoon Winnie(9711) and Bilis(0010) underwent extratropical transition (ET) over Chinese mainland but only Winnie re-intensified as an extratropical system. Their ET processes were examined and compared in terms of the Moisture Potential Vorticity u ( m P ) perspective. Results show that Winnie enters mid-latitude zone and couples with an upper trough during its ET. Bilis also approaches an upper trough but decouples with it. ET re-intensification is related to the interactions between typhoon remnant circulation and lower layer front zone and m P anomaly downward transported from the upper troposphere. Especially, slantwise vorticity development caused by the increase of moist baroclinicity is an main factor responsible for Winnie re-intensification. However, there is no frontal zone occurring in Bilis鈥檚 remnant circulation and downward transportation of the upper m P anomaly is weak.
    Mak M., 1991: Dynamics of an atmospheric blocking as deduced from its local energetics. Quart. J. Roy. Meteor. Soc., 117, 477- 493.10.1002/qj.497117499040927dd994cf529fb0205415a142a6536http%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1002%2Fqj.49711749904%2Fabstracthttp://onlinelibrary.wiley.com/doi/10.1002/qj.49711749904/abstractAbstract This paper reports an analysis of the roles of temporal scale interactions (seasonal, intraseasonal and high-frequency components) in the generation and maintenance of a pronounced block over the North Atlantic that lasted for three weeks in February 1983. This blocking disturbance had a distinct dipole structure initially oriented in an east-west direction downstream from a strongly diffluent south-west-north-east oriented seasonal jet. Although it has an equivalent barotropic structure, there is a well-defined vertical velocity field with ascending (descending) motion on its western (eastern) flank. As it develops, matures and decays, it rotates systematically in a clockwise direction. The contributions from the various temporal scale interactions to the episodal average local energetics of this block are evaluated. There are five comparably important processes controlling the intensity, configuration, and evolution of the block. The synoptic eddy-straining mechanism proposed by Shutts is manifested in three energetics terms, of which one is found to be particularly large. The blocking disturbance also barotropically extracts kinetic energy at a significant rate from the seasonal diffluent jet under the influence of the latter's strong deformation field. The pressure work process, the baroclinic conversion process, and above all, the nonlinear dynamics of the blocking disturbance itself are quantitatively important in redistributing the energy within the blocking region. The effects of the diabatic and subgridscale processes are found, as residues, to be substantially dissipative on the block.
    Mak M., M. Cai, 1989: Local barotropic instability. J. Atmos. Sci., 46, 3289- 3311.36098dea-8394-4887-aa29-799495879cd309e5f4fa74510a3ccfe700282d1c88e1http%3A%2F%2Fwww.researchgate.net%2Fpublication%2F234539544_Local_Barotropic_Instabilityrefpaperuri:(f5e9dd68688a0438f76ccae9f19f75bf)http://www.researchgate.net/publication/234539544_Local_Barotropic_Instability
    Meng W. G., A. Y. Wang, J. N. Li, R. Q. Feng, and E. B. Hou, 2004: Moist potential vorticity analysis of the heavy rainfall and mesoscale convective systems in South China. Chinese J. Atmos. Sci., 28( 3), 330- 341. (in Chinese)6b513eff-d3b2-42e0-8f02-c5c1155cde9f0215ac06e8a96bc7b12cda39cd0ae104http%3A%2F%2Fen.cnki.com.cn%2FArticle_en%2FCJFDTOTAL-DQXK200403001.htmrefpaperuri:(871ca2f101da3d6a8f547efdaed0e607)http://en.cnki.com.cn/Article_en/CJFDTOTAL-DQXK200403001.htmBy using the MM5 model outputs of a successful numerical simulation on a South China heavy rainfall event and the mesoscale convective system (MCS) occurred during 23锝24 May 1998, the development of the heavy rainfall and MCS have been investigated in terms of moist potential vorticity principle and slantwise vorticity development theory The results show that, on the slantwise moist isentropic surface, areas with high pressure and positive moist potential vorticity are favorable for the development of heavy rainfall and MCS As the cold air slide down along the moist isentropic surface and confluence with the slantwise upward motion warm and moist air with high convective available potential energy values, both of them experience a stability decreasing process, and lead to cyclonic vorticity development Over deep convection area, the atmosphere exhibits the signature of conditional symmetric instability, and MCS is characterized by slantwise upward motion As for the slantwise of moist isentropic surface, the increase in vertical shear of horizontal wind or enhancement in moist baroclinity also resulted in the increase of vertical vorticity and the development of MCS Finally in the article, a physical conceptual model about the development of heavy rainfall and MCS over South China was presented
    Moon Y., D. S. Nolan, 2015: Spiral rainbands in a numerical simulation of Hurricane Bill (2009). Part II: Propagation of inner rainbands. J. Atmos. Sci., 72, 191- 215.10.1175/JAS-D-14-0056.1e16bc3ebe74d8ba64a7f93cdda74dbachttp%3A%2F%2Fwww.researchgate.net%2Fpublication%2F273193041_Spiral_Rainbands_in_a_Numerical_Simulation_of_Hurricane_Bill_%282009%29._Part_II_Propagation_of_Inner_Rainbands%3Fev%3Dauth_pubhttp://www.researchgate.net/publication/273193041_Spiral_Rainbands_in_a_Numerical_Simulation_of_Hurricane_Bill_(2009)._Part_II_Propagation_of_Inner_Rainbands?ev=auth_pubAbstract This is the second part of a study that examines spiral rainbands in a numerical simulation of Hurricane Bill (2009). This paper evaluates whether the propagation of inner rainbands in the Hurricane Bill simulation is consistent with previously proposed hypotheses. Results indicate that the propagation of inner rainbands is not consistent with gravity waves, vortex Rossby waves, or squall lines. An alternative hypothesis is offered, arguing that inner rainbands are simply convective clouds that are advected by the rapidly rotating tropical cyclone wind field while being deformed into spiral shapes. A summary and a discussion of the results of both Parts I and II are provided.
    Petterssen S., 1956: Weather Analysis and Forecasting Vol. I: Motion and Motion Systems. McGraw-Hill,428 pp.
    Renfrew I. A., A. J. Thorpe, and C. H. Bishop, 1997: The role of the environmental flow in the development of secondary frontal cyclones. J. Atmos. Sci. , 123, 1653- 1675.10.1002/qj.49712354210f0c6ccaa9e8c181df8422748ecad8718http%3A%2F%2Fonlinelibrary.wiley.com%2Fdoi%2F10.1002%2Fqj.49712354210%2Fabstracthttp://onlinelibrary.wiley.com/doi/10.1002/qj.49712354210/abstractAbstract The impact of the environmental flow on the development of secondary frontal cyclones is investigated. Several case-studies are examined as examples of secondary frontal-cyclone events observed in the North Atlantic-western Europe sector. A simple measure of growth is defined to chart their development. The vorticity attribution technique of Bishop is utilized to calculate the action of the large-scale (environmental) flow on the fronts. In particular the environmental along-front stretching—shown to be important in theoretical models of frontal instabilities—is calculated. The role of the environmental deformation appears to be crucial: as part of a baroclinic life cycle, stretching deformation acts to build up a front but suppress along-front waves; if the stretching rate diminishes, barotropic instabilities may then break out. Diagnostics are examined to try to ascertain the growth mechanisms at work in each frontal-cyclone case. A range of values for the commonly prescribed deformation-frontogenesis and shearing-frontogenesis parameters are calculated.
    Rivière G., A. Joly, 2006a: Role of the low-frequency deformation field on the explosive growth of extratropical cyclones at the jet exit. Part I: Barotropic critical region. J. Atmos. Sci., 63, 1965- 1981.
    Rivière G., A. Joly, 2006b: Role of the low-frequency deformation field on the explosive growth of extratropical cyclones at the jet exit. Part II: Baroclinic critical region. J. Atmos. Sci., 63, 1982- 2006.10.1175/JAS3729.107ee4af4-0db8-4eb6-89e1-6aa204ee50caa8d779b1335da7c9de6dcdc5b9d7c93ahttp%3A%2F%2Fwww.researchgate.net%2Fpublication%2F259324283_Role_of_the_Low-Frequency_Deformation_Field_on_the_Explosive_Growth_of_Extratropical_Cyclones_at_the_Jet_Exit._Part_I_Barotropic_Critical_Regionrefpaperuri:(19053d567a8c753a6efba2a26b70d217)http://www.researchgate.net/publication/259324283_Role_of_the_Low-Frequency_Deformation_Field_on_the_Explosive_Growth_of_Extratropical_Cyclones_at_the_Jet_Exit._Part_I_Barotropic_Critical_RegionBy using new theoretical results on perturbation growth in spatially and temporally complex quasigeostrophic flows, this paper investigates the role of the large-scale deformation field on extratropical cyclones and especially on their explosive growth in the jet-exit region. Theoretical ideas are tested by decomposing the atmospheric flow into a high- and a low-frequency part and by analyzing four-dimensional variational data assimilation (4DVAR) reanalysis data of the Fronts and Atlantic Storm-Track Experiment (FASTEX) during February 1997 as well as reanalysis data for the end of December 1999. Regions where the low-frequency deformation magnitude is greater than the absolute value of the low-frequency vorticity are shown to correspond to regions where synoptic disturbances at the same level tend to be located. These regions in the upper troposphere are intrinsically related to the horizontal inhomogeneities of the low-frequency large-scale upper-tropospheric jet but cannot be detected by looking separately at the deformation or vorticity. Transitions from one such large-scale region to the next furthermore can be accompanied by a sudden change of the dilatation axes orientation: this combination defines a barotropic critical region (BtCR). Reasons why a BtCR is a specific place where barotropic development is likely to occur are exposed. Two very differently located BtCR regions in two apparently similar zonal-like weather regimes are shown to be the preferred regions where synoptic eddies tend to cross the jet from the south to the north. BtCRs are also special regions where constructive association between barotropic and baroclinic processes is favored, indeed constrained to cooperate. This is illustrated through the detailed analysis of the last growth stage of Intensive Observation Period 17 (IOP17) of FASTEX. It happens precisely around a BtCR area located in the jet-exit region. Two processes explain this IOP17 development; one involves the barotropic generation rate resulting from the low crossing the BtCR and the other one is baroclinic interaction, which is strongly maintained far away from the baroclinicity maximum because of the new favorable baroclinic configuration resulting from the first process.
    Rozoff C. M., W. H. Schubert, B. D. McNoldy, and J. P. Kossin, 2006: Rapid filamentation zones in intense tropical cyclones. J. Atmos. Sci., 63, 325- 340.10.1175/JAS3595.1f5d2d9babcf3f8effffcd16c72341b6bhttp%3A%2F%2Fwww.researchgate.net%2Fpublication%2F253655601_Rapid_Filamentation_Zones_in_Intense_Tropical_Cycloneshttp://www.researchgate.net/publication/253655601_Rapid_Filamentation_Zones_in_Intense_Tropical_CyclonesAbstract Intense tropical cyclones often possess relatively little convection around their cores. In radar composites, this surrounding region is usually echo-free or contains light stratiform precipitation. While subsidence is typically quite pronounced in this region, it is not the only mechanism suppressing convection. Another possible mechanism leading to weak-echo moats is presented in this paper. The basic idea is that the strain-dominated flow surrounding an intense vortex core creates an unfavorable environment for sustained deep, moist convection. Strain-dominated regions of a tropical cyclone can be distinguished from rotation-dominated regions by the sign of S 2 1 + S 2 2 61 ζ 2 , where S 1 = u x 61 υ y and S 2 = υ x + u y are the rates of strain and ζ = υ x 61 u y is the relative vorticity. Within the radius of maximum tangential wind, the flow tends to be rotation-dominated ( ζ 2 > S 2 1 + S 2 2 ), so that coherent structures, such as mesovortices, can survive for long periods of time. Outside the radius of maximum tangential wind, the flow tends to be strain-dominated ( S 2 1 + S 2 2 > ζ 2 ), resulting in filaments of anomalous vorticity. In the regions of strain-dominated flow the filamentation time is defined as τ fil = 2( S 2 1 + S 2 2 61 ζ 2 ) 611/2 . In a tropical cyclone, an approximately 30-km-wide annular region can exist just outside the radius of maximum tangential wind, where τ fil is less than 30 min and even as small as 5 min. This region is defined as the rapid filamentation zone. Since the time scale for deep moist convective overturning is approximately 30 min, deep convection can be significantly distorted and even suppressed in the rapid filamentation zone. A nondivergent barotropic model illustrates the effects of rapid filamentation zones in category 1–5 hurricanes and demonstrates the evolution of such zones during binary vortex interaction and mesovortex formation from a thin annular ring of enhanced vorticity.
    Shutts G. J., 1983: The propagation of eddies in diffluent jet-streams: Eddy vorticity forcing of "blocking" flow fields. Quart. J. Roy. Meteor. Soc., 109, 737- 761.be53ba0f-2a54-4b1e-9a7c-78d761405086e160dd757d20139de33816a5fbc74acahttp%3A%2F%2Fciteseer.ist.psu.edu%2Fshowciting%3Fcid%3D4439989refpaperuri:(405d7dfc484c19a6d8e0d3522c43d469)http://citeseer.ist.psu.edu/showciting?cid=4439989
    Spensberger C., T. Spengler, 2014: A new look at deformation as a diagnostic for large-scale flow. J. Atmos. Sci., 71, 4221- 4234.4718137e2807d486774cd37cc3c98704http%3A%2F%2Fwww.researchgate.net%2Fpublication%2F270719738_A_New_Look_at_Deformation_as_a_Diagnostic_for_Large-Scale_Flowhttp://www.researchgate.net/publication/270719738_A_New_Look_at_Deformation_as_a_Diagnostic_for_Large-Scale_Flow
    Thomas L. N., 2012: On the effects of frontogenetic strain on symmetric instability and inertia-gravity waves. J. Fluid Mech., 711, 620- 640.10.1017/jfm.2012.4168d3dea0f-9168-4113-b9c7-490328ccfc63d0c7da0867971f71e91c62ac76317647http%3A%2F%2Fjournals.cambridge.org%2Fabstract_S0022112012004168refpaperuri:(8e2b9f0b17e33a17cc42be2fd240abc8)http://journals.cambridge.org/abstract_S0022112012004168The dynamics of symmetric instability and two-dimensional inertiagravity waves, depending on the sign of the Ertel potential vorticity and the magnitude of the Richardson number of the geostrophic flow. The kinetic energy (KE) of both types of motion is suppressed by frontogenetic strain due to the vertical shear in the ageostrophic circulation. This is because the perturbation streamlines tilt with the ageostrophic shear causing the disturbances to lose KE via shear production. The effect can completely dampen symmetric instability for sufficiently strong strain even though the source of KE for the instability (the vertical shear in the geostrophic flow) increases with time. Inertiagravity waves play a catalytic role in loss of balance. Given the large amount of KE in low-frequency inertia鈥揼ravity waves and the ubiquitous combination of strain and baroclinic geostrophic currents in the ocean, it is estimated that this mechanism could play a significant role in the removal of KE from both the internal wave and mesoscale eddy fields.
    Wang J. Z., S. F. Ma, and Y. H. Ding, 1996: Application of potential vorticity theory to analysis of formative mechanism of torrential rain. Quarterly Journal of Applied Meteorology, 7, 19- 27.7747c6d93c9a52d7285130a0e58538d5http%3A%2F%2Fqk.cams.cma.gov.cn%2Fjams%2Fch%2Freader%2Fview_abstract.aspx%3Ffile_no%3D19960103http://qk.cams.cma.gov.cn/jams/ch/reader/view_abstract.aspx?file_no=19960103应用位涡理论,对1991年江淮地区一次特大暴雨过程中位涡及其有关物理量的分布特征进行了分析。结果表明:强降水总是落在干位涡比较小的地方和湿位涡负中心暖气流一侧,它和相对湿位涡的关系更直接。湿位涡斜压部分可清楚地反映湿斜压性对对流不稳定系统所起的作用。大气在湿位涡值比较小的区域对锋生强迫有更强的响应。
    Wang X. B., R. S. Wu, 2001: The development of symmetric disturbance superposed on baroclinic frontal zone under the action of deformation field. Acta Meteorologica Sinica, 15( 4), 420- 435.10.3969/j.issn.0894-0525.2001.04.004bad12f8ce3578eb4fa6d9498164ee8dbhttp%3A%2F%2Fwww.cqvip.com%2FMain%2FDetail.aspx%3Fid%3D1001466128http://d.wanfangdata.com.cn/Periodical_qxxb-e200104004.aspx正 The development of symmetric disturbance superposed on the background field of Hoskins-Bretherton (1972) frontogenesis model is investigated by means of WKBJ approach,It is found thatthe forcing of large-scale deformation,the frontal circulation and the spatial-temporal variations ofstability para
    Wang Y. Q., 2008: Rapid filamentation zone in a numerically simulated tropical cyclone. J. Atmos. Sci., 65, 1158- 1181.10.1175/2007JAS2426.137786838-0861-40b7-935f-d2b7b8d82fcc04ad4bbb83fd98f120f9f748be27c0achttp%3A%2F%2Fwww.researchgate.net%2Fpublication%2F249610529_Rapid_Filamentation_Zone_in_a_Numerically_Simulated_Tropical_Cyclonerefpaperuri:(1fd3851a80d683c1fcb8d554a2134afe)http://www.researchgate.net/publication/249610529_Rapid_Filamentation_Zone_in_a_Numerically_Simulated_Tropical_CycloneAbstract In a recent study, Rozoff et al. proposed a possible mechanism to explain the formation and maintenance of the weak-echo annulus (or moat) outside of the primary eyewall of a tropical cyclone observed in radar images. By this mechanism, the moat is determined to be a region of the strain-dominated flow outside of the radius of maximum wind in which essentially all fields are filamented and deep convection is hypothesized to be highly distorted and even suppressed. This strain-dominated region is defined as the rapid filamentation zone wherein the filamentation time is shorter than the overturning time of deep convection. An attempt has been made in this study to test the hypothesis in a full-physics tropical cyclone model under idealized conditions and to extend the concept to the study of the inner-core dynamics of tropical cyclones. The foci of this paper are the evolution of the rapid filamentation zone during the storm intensification, the potential roles of rapid filamentation in the organization of inner spiral rainbands, and the damping of high azimuthal wavenumber asymmetries in the tropical cyclone inner core. The presented results show that instead of suppressing deep convection, the strain flow in the rapid filamentation zone outside the elevated potential vorticity core provides a favorable environment for the organized inner spiral rainbands, which generally have time scales of several hours, much longer than the typical overturning time scale of individual convective clouds. Although the moat in the simulated tropical cyclone is located in the rapid filamentation zone, it is mainly controlled by the subsidence associated with the overturning flow from eyewall convection and downdrafts from the anvil stratiform precipitation outside of the eyewall. It is thus suggested that rapid filamentation is likely to play a secondary role in the formation of the moat in tropical cyclones. Although the deformation field is determined primarily by the structure of the tropical cyclone, it can have a considerable effect on the evolution of the storm. Because of strong straining deformation, asymmetries with azimuthal wavenumber >4 are found to be damped effectively in the rapid filamentation zone. The filamentation time thus provides a quantitative measure of the stabilization and axisymmetrization of high-wavenumber asymmetries in the inner core by shearing deformation and filamentation.
    Wang Z. Q., W. J. Zhu, and A. M. Duan, 2010: A case study of snowstorm in Tibetan Plateau induced by Bay of Bengal storm: Based on the theory of slantwise vorticity development. Plateau Meteorology, 29( 3), 703- 711. (in Chinese) 0e05424a-7ea2-4657-8eab-69101305b34amag484262010293703<FONT face=Verdana>Based on the diagnosis of moist potential vorticity (MPV) and the theory of slantwise vorticity development (SVD), a case study of snowstorm in the Tibetan Plateau induced mainly by Bay of Bengal storm in November 2007 was analyzed. The results show that the interaction of the storm′s spiral cloud band and the down\|sliding dry cold air (positive MPV1 band) over the steep southern fringe of the plateau is the main reason for this synoptic process. Owing to the dense and obvious slantwise isentropic surface, the contribution of negative MPV2 to moist potential vorticity (MPV) is more than MPV1 over the snowfall area where atmosphere is weak stable. It further leads to the sharp growth of SVD. Moreover, negative MPV makes a favorable background of the generation of conditional symmetrical instability. The conditional symmetrical instability is beneficial tothe development of slantwise vorticity, and it is a possible important mechanism for this snowstorm.</FONT>
    Weiss J., 1991: The dynamics of enstrophy transfer in two-dimensional hydrodynamics. Physica D: Nonlinear Phenomena, 48, 273- 294.10.1016/0167-2789(91)90088-Qc0b3962907582201d55af1945172624ahttp%3A%2F%2Fwww.sciencedirect.com%2Fscience%2Farticle%2Fpii%2F016727899190088Qhttp://www.sciencedirect.com/science/article/pii/016727899190088QIn this paper the qualitative properties of an inviscid, incompressible, two-dimensional fluid are examined. Starting from the equations of motion we derive a series of equations that govern the behavior of the spatial gradients of the vorticity scalar. The growth of these gradients is related to the transfer of enstrophy (integral of squared vorticity) to the small scales of the fluid motion.
    Whitaker J. S., R. M. Dole, 1995: Organization of storm tracks in zonally varying flows. J. Atmos. Sci., 52( 8), 1178- 1191.10.1175/1520-0469(1995)0522.0.CO;2015cdbefd6de48d6449fdbb29a21e90chttp%3A%2F%2Fwww.researchgate.net%2Fpublication%2F249608473_Organization_of_Storm_Tracks_in_Zonally_Varying_Flowshttp://www.researchgate.net/publication/249608473_Organization_of_Storm_Tracks_in_Zonally_Varying_FlowsAbstract A simple two-layer quasigeostrophic model is employed to investigate the sensitivity of storm tracks to changes in an externally imposed, zonally varying large-scale flow. Zonally asymmetric temperature and horizontal deformation fields are varied systematically in order to compare the effects of baroclinicity and horizontal deformation on storm track dynamics. The sensitivity of the storm tracks to uniform barotropic zonal flows is also examined. The results show two competing processes for storm track organization, one associated with a local maximum in baroclinicity and the other with a local minimum in horizontal deformation. When the equilibrium state consists of a zonally symmetric temperature field and a barotropic stationary wave, the maximum in synoptic-scale transient eddy energy (storm track) is located in the entrance region of the upper jet just downstream of the point of minimum horizontal deformation. As zonal variations in baroclinicity become large (keeping the upper-layer horizontal deformation constant), the storm track shifts to the jet exit region just downstream of the point of maximum baroclinicity. For flows intermediate between the above cases,that is, having weaker zonal variations in baroclinicity and the same upper-layer deformation, two storm track maxima appear, one located in the jet entrance and the other in the jet exit region. The results also indicate that the storm tracks are sensitive to changes in a uniform barotropic zonal flow. The presence of a uniform westerly flow extends the storm track and strengthens eddy activity, while the addition of a uniform easterly flow shortens the storm track and dramatically weakens eddy activity. The changes in the magnitudes of eddy activity appear related to differences in the efficiency of nonlinear barotropic decay processes in weakening the eddies in the jet exit region. Sensitivities of the location of the storm tracks to changes in large-scale flow parameters are well captured by linear calculations, although sensitivities of the strength of the storm tracks are not. For sufficiently strong zonal variations in baroclinicity, two coherent modes of low-frequency variability develop. They are characterized synoptically by 1) a meridional shift, and 2) an extension/contraction as well as a modulation in the strength of the upper-layer jet and storm track.
    Wu G. X., 2001: Comparison between the complete-form vorticity equation and the traditional vorticity equation. Acta Meteorologica Sinica, 59( 4), 385- 392. (in Chinese)10.11676/qxxb2001.042ad52a165-6d8c-4ae8-83c0-c5d8512b11cd5584200142After a brief review on the deductions of the complete-form vorticity equation and traditional vorticity equation,comparisons between the to equations are made.Particular attentions are paid to the new physical meanings contained in the complete-form vorticity equation.It is revealed the mechanism for vorticity development due to the internal forcing associated with the changes in static stability and baroclinicity (frontogenesis or frontolysis),as well as the external forcing associated with frictional dissipation and diabatic heating.Finaly,the application prospects of the complete-form vorticity equation in weather and climate studies are discussed.
    Wu G. X., Y. P. Cai, 1997: Vertical wind shear and down-sliding slantwise vorticity development. Scientia Atmospherica Sinica, 21( 3), 273- 282. (in Chinese)3264bb0d0cc3320d8fc739bb4fbf22e7http%3A%2F%2Fen.cnki.com.cn%2FArticle_en%2FCJFDTOTAL-DQXK703.002.htmhttp://en.cnki.com.cn/Article_en/CJFDTOTAL-DQXK703.002.htmBased upon the conservation property of moist potential vorticity ( P m ) of an adiabatic, frictionless, and saturated atmosphere, the development of vertical vorticity in a moist baroclinic process was discussed. When moist isentropic surfaces are tilted, the application of the traditional “isentropic potential vorticity” (IPV) analysis is limited. A theory of slantwise vorticity development was then developed to investigate the vorticity intensification from a Lagrangian point of view. It was shown that in the area between the south of monsoon front and the north of warm and moist air mass, moist isentropic surfaces are stiff. This area then becomes a favorable region for the development of cyclone and torrential rain. The necessary condition and sufficient condition for slantwise vorticity development are discussed. It is proved that in a convectively unstable and saturated atmosphere, the occurence of slantwise vorticity development must be accompanied by the existence of a low level jet. Application of this theory to a case analyses of typical monsoonal torrential rain shows that the P m analysis, especially the analysis of P m1 (=- g(f+ζ p)e/) and P m2 (=- g×/· pθ e ) at isobaric surfaces in the lower troposphere, is very effective in identifying the occurence of torrential rain, and may be used as a powerful tool for the diagnosis and prediction of torrential rain.
    Wu G. X., H. Z. Liu, 1999: Complete form of vertical vorticity tendency equation and slantwise vorticity development. Acta Meteorologica, 57( 1), 1- 15. (in Chinese)3a4ae8ab7aaa06a6a60283e350722c66http%3A%2F%2Fwww.cmsjournal.net%2Fqxxben%2Fch%2Freader%2Fview_abstract.aspx%3Ffile_no%3D19990101%26flag%3D1http://www.cmsjournal.net/qxxben/ch/reader/view_abstract.aspx?file_no=19990101&amp;flag=1
    Wu G. X., Y. P. Cai, and X. J. Tang, 1995: Moist potential vorticity and slantwise vorticity development. Acta Meteorologica Sinica, 53( 4), 387- 405. (in Chinese)10.11676/qxxb1995.0459cdfc5b2-5e47-4e5f-bd81-cc3a8eee812e55841995467ee25f9c92a212174f27e57fc92a5cdchttp%3A%2F%2Fen.cnki.com.cn%2FArticle_en%2FCJFDTOTAL-QXXB504.001.htmhttp://en.cnki.com.cn/Article_en/CJFDTOTAL-QXXB504.001.htmAn accurate form of the moist potential vorticity (MPV) equation was deduced from a complete set of primitive equations system. It was shown that motion in a saturated atmosphere without diabatic heating and frictional dissipation conserves moist potential vorticity.This property was then used to investigate the development of vertical vorticity in moist baroclinic processes. Results show that in the frame work of moist isentropic coordinate, vorticity development can result from reduction of convective stability, or convergence. or latent heat release in isentropic surfaces. However, the application of the usual analysis of moist isentropic potential vorticity is limitted due to the declination of moist isentropic surfaces, and a theory of slantwise vorticity development based on Z-coordinate and P-coordinate was then proposed. According to this theory, whether the atmosphere is moist symetrically stable or unstable. or convectively stable or unstable. so long as the moist insentropic surface is slantwise the reduction of convective stability, the increase of the vertical shear of horizontal wind or moist baroclinity can result in the increase of vertical vorticity. The larger the declination of the moist isentropes, the more vigorous the deveolopment of vertical vorticity. In a region with a monsoonal front to the north and warm and moist air to the south, or bye the north of front the moist isentropes are very stiff. This is the region most favorable for the development of vorticities and formation of torrential rain.For a case of persistent torrential rain occuriny in the middle and lower reaches of the Yangtze and Huai River in June 12-15, 1991. moist potential vorticity analysis, especially the isobaric analysis of its vertical and horizontal compoments, i e. MPVI and MPV2 respectively, is effective for indentifying synoptic systems not only in middle and high latitudes. but also in the low latitudes and in the lower troposphere. It can serve as a powerful tool for the diagnosis and prediction of torrential rain.
    Wu G. X., Y. J. Zheng, and Y. M. Liu, 2013: Dynamical and thermal problems in vortex development and movement. Part II: Generalized slantwise vorticity development. Acta Meteorologica Sinica, 27( 1), 15- 25.10.1007/s13351-013-0102-2168bf3d2fb3e2969bcd93f84738d9f6ahttp%3A%2F%2Fwww.cnki.com.cn%2FArticle%2FCJFDTotal-QXXW201301002.htmhttp://d.wanfangdata.com.cn/Periodical_qxxb-e201301002.aspxThe development of vertical vorticity under adiabatic condition is investigated by virtue of the view of potential vorticity and potential temperature (PV-胃) and from a Lagrangian perspective. A new concept of generalized slantwise vorticity development (GSVD) is introduced for adiabatic condition. The GSVD is a coordinate independent framework of vorticity development (VD), which includes slantwise vorticity development (SVD) when a particle is sliding down the concave slope or up the convex slope of a sharply tilting isentropic surface under stable or unstable condition. The SVD is a special VD for studying the severe weather systems with rapid development of vertical vorticity. In addition, the GSVD clarifies VD and SVD. The criteria for VD and SVD demonstrate that the demand for SVD is much more restricted than the demand for VD. When an air parcel is moving down the concave slope or up the convex slope of a sharply tilting isentropic surface in a stable stratified atmosphere with its stability decreasing, or in an unstable atmosphere with its stability increasing, i.e., its stability 胃 z approaches zero, its vertical vorticity can develop rapidly if its C D is decreasing. The theoretical results are employed to analyze a Tibetan Plateau (TP) vortex (TPV), which appeared over the TP, then slid down and moved eastward in late July 2008, resulting in heavy rainfall in Sichuan Province and along the middle and lower reaches of the Yangtze River. The change of PV 2 contributed to the intensification of the TPV from 0000 to 0600 UTC 22 July 2008 when it slid upward on the upslope of the northeastern edge of the Sichuan basin, since the changes in both horizontal vorticity 畏 s and baroclinity 胃 s have positive effects on the development of vertical vorticity. At 0600 UTC 22 July 2008, the criterion for SVD at 300 K isentropic surface is satisfied, meaning that SVD occurred and contributed significantly to the development of vertical vorticity. The appearance of the stronger signals concerning the VD and SVD surrounding the vortex indicates that the GSVD concept can serve as a useful tool for diagnosing the development of weather systems.
    Yang S., S. T. Gao, and C. G. Lu, 2014: A generalized frontogenesis function and its application. Adv. Atmos. Sci.,31(5), 1065-1078, doi: 10.1007/s00376-014-3228-y.
    Yang S., S. T. Gao, and C. G. Lu, 2015: Investigation of the Mei-yu front using a new deformation frontogenesis function. Adv. Atmos. Sci.,32(5), 635-647, doi: 10.1007/s00376-014-4147-7.
    Yu H., G. X. Wu, 2001: Moist baroclinicity and abrupt intensity change of tropical cyclone. Acta Meteorologica Sinica, 59( 4), 440- 449 (in Chinese).
    Zheng Y. J., G. X. Wu, and Y. M. Liu, 2013: Dynamical and thermal problems in vortex development and movement. Part I: A PV-Q view. Acta Meteorologica Sinica, 27( 2), 1- 14.
  • [1] YANG Shuai, GAO Shouting, Chungu LU, 2015: Investigation of the Mei-yu Front Using a New Deformation Frontogenesis Function, ADVANCES IN ATMOSPHERIC SCIENCES, 32, 635-647.  doi: 10.1007/s00376-014-4147-7
    [2] GAO Shouting, YANG Shuai, XUE Ming, CUI Chunguang, 2008: Total Deformation and its Role in Heavy Precipitation Events Associated with Deformation-Dominant Flow Patterns, ADVANCES IN ATMOSPHERIC SCIENCES, 25, 11-23.  doi: 10.1007/s00376-008-0011-y
    [3] 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
    [4] YANG Shuai, GAO Shouting, LU Chungu, 2014: A Generalized Frontogenesis Function and Its Application, ADVANCES IN ATMOSPHERIC SCIENCES, 31, 1065-1078.  doi: 10.1007/s00376-014-3228-y
    [5] REN Rongcai, Ming CAI, 2006: Polar Vortex Oscillation Viewed in an Isentropic Potential Vorticity Coordinate, ADVANCES IN ATMOSPHERIC SCIENCES, 23, 884-900.  doi: 10.1007/s00376-006-0884-6
    [6] CHEN Min, ZHENG Yongguang, 2004: Vorticity Budget Investigation of a Simulated Long-Lived Mesoscale Vortex in South China, ADVANCES IN ATMOSPHERIC SCIENCES, 21, 928-940.  doi: 10.1007/BF02915595
    [7] Wei TAO, Linlin ZHENG, Ying HAO, Gaoping LIU, 2023: An Extreme Gale Event in East China under the Arctic Potential Vorticity Anomaly through the Northeast China Cold Vortex, ADVANCES IN ATMOSPHERIC SCIENCES, 40, 2169-2182.  doi: 10.1007/s00376-023-2255-y
    [8] Xiuping YAO, Qin ZHANG, Xiao ZHANG, 2020: Potential Vorticity Diagnostic Analysis on the Impact of the Easterlies Vortex on the Short-term Movement of the Subtropical Anticyclone over the Western Pacific in the Mei-yu Period, ADVANCES IN ATMOSPHERIC SCIENCES, 37, 1019-1031.  doi: 10.1007/s00376-020-9271-y
    [9] Na LI, Lingkun RAN, Linna ZHANG, Shouting GAO, 2017: Potential Deformation and Its Application to the Diagnosis of Heavy Precipitation in Mesoscale Convective Systems, ADVANCES IN ATMOSPHERIC SCIENCES, 34, 894-908.  doi: 10.1007/s00376-017-6282-4
    [10] Gao Shouting, Lei Ting, 2000: Streamwise Vorticity Equation, ADVANCES IN ATMOSPHERIC SCIENCES, 17, 339-347.  doi: 10.1007/s00376-000-0027-4
    [11] Brian HOSKINS, 2015: Potential Vorticity and the PV Perspective, ADVANCES IN ATMOSPHERIC SCIENCES, 32, 2-9.  doi: 10.1007/s00376-014-0007-8
    [12] Zuohao Cao, 1999: Dynamics of Absolute Vorticity in the Boussinesq Fluid, ADVANCES IN ATMOSPHERIC SCIENCES, 16, 482-486.  doi: 10.1007/s00376-999-0025-0
    [13] Zuohao CAO, Da-Lin ZHANG, 2004: Tracking Surface Cyclones with Moist Potential Vorticity, ADVANCES IN ATMOSPHERIC SCIENCES, 21, 830-835.  doi: 10.1007/BF02916379
    [14] Fang Juan, Wu Rongsheng, 1998: Influences of Vorticity Source and Momentum Source on Atmospheric Circulation, ADVANCES IN ATMOSPHERIC SCIENCES, 15, 41-46.  doi: 10.1007/s00376-998-0016-6
    [15] Chanh Q. KIEU, Da-Lin ZHANG, 2012: Is the Isentropic Surface Always Impermeable to the Potential Vorticity Substance?, ADVANCES IN ATMOSPHERIC SCIENCES, 29, 29-35.  doi: 10.1007/s00376-011-0227-0
    [16] CUI Xiaopeng, GAO Shouting, WU Guoxiong, 2003: Up-Sliding Slantwise Vorticity Development and the Complete Vorticity Equation with Mass Forcing, ADVANCES IN ATMOSPHERIC SCIENCES, 20, 825-836.  doi: 10.1007/BF02915408
    [17] Gang LI, Daoyong YANG, Xiaohua JIANG, Jing PAN, Yanke TAN, 2017: Diagnosis of Moist Vorticity and Moist Divergence for a Heavy Precipitation Event in Southwestern China, ADVANCES IN ATMOSPHERIC SCIENCES, 34, 88-100.  doi: 10.1007/s00376-016-6124-9
    [18] Zuohao CAO, Da-Lin ZHANG, 2005: Sensitivity of Cyclone Tracks to the Initial Moisture Distribution: A Moist Potential Vorticity Perspective, ADVANCES IN ATMOSPHERIC SCIENCES, 22, 807-820.  doi: 10.1007/BF02918681
    [19] Zuohao Cao, G.W.K. Moore, 1998: A Diagnostic Study of Moist Potential Vorticity Generation in an Extratropical Cyclone, ADVANCES IN ATMOSPHERIC SCIENCES, 15, 152-166.  doi: 10.1007/s00376-998-0036-2
    [20] Olivia MARTIUS, Cornelia SCHWIERZ, Michael SPRENGER, 2008: Dynamical Tropopause Variability and Potential Vorticity Streamers in the Northern Hemisphere ---A Climatological Analysis, ADVANCES IN ATMOSPHERIC SCIENCES, 25, 367-380.  doi: 10.1007/s00376-008-0367-z

Get Citation+

Export:  

Share Article

Manuscript History

Manuscript received: 24 March 2015
Manuscript revised: 07 July 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

The Impact of Deformation on Vortex Development in a Baroclinic Moist Atmosphere

  • 1. Key Laboratory of Cloud-Precipitation Physics and Severe Storms, Institute of Atmospheric Physics, Chinese Academy of Sciences, Beijing 100029
  • 2. State Key Laboratory of Severe Weather, Chinese Academy of Meteorological Sciences, Beijing 100081

Abstract: A mathematical relation between deformation and vertical vorticity tendency is built by introducing the frontogenesis function and the complete vertical vorticity equation, which is derived by virtue of moist potential vorticity. From the mathematical relation, it is shown that properly configured atmospheric conditions can make deformation exert a positive contribution to vortex development at rates comparable to other favorable factors. The effect of deformation on vortex development is not only related to the deformation itself, but also depends on the current thermodynamic and dynamic structures of the atmosphere, such as the convective stability, moist baroclinicity and vertical wind shear (or horizontal vorticity). A diagnostic study of a heavy-rainfall case that occurred during 20-22 July 2012 shows that deformation has the most remarkable effect on the increase in vertical vorticity during the rapid development stage of the low vortex during its whole life cycle. This feature is mainly due to the existence of an approximate neutral layer (about 700 hPa) in the atmosphere where the convective stability tends to be zero. The neutral layer makes the effect of deformation on the vertical vorticity increase significantly during the vortex development stage, and thus drives the vertical vorticity to increase.

1. Introduction
  • In meteorology, deformation is mostly applied to the frontogenesis process (Petterssen, 1956; Yang et al., 2014; Yang et al., 2015). However, many observations and a number of clear and simple theories show that deformation also plays important roles in other physical processes, such as storm-track dynamics, the formation and maintenance of the moat structure in tropical cyclones, symmetric instability, blocking onset, and so on (e.g., Whitaker and Dole, 1995; Elhmaidi et al., 2004; Rivière and Joly, 2006a, b; Gao et al., 2008; Wang, 2008; Thomas, 2012; Moon and Nolan, 2015). Based on the traditional exponential growth problem of normal modes (Charney, 1947; Eady, 1949; Kuo, 1949), (Farrell, 1989) demonstrated that, although the deformational flow does not support any exponentially growing solution, disturbance embedded in the deformational flow undergoes transient development. (Mak and Cai, 1989) obtained a general condition for the stability property of a perturbation embedded in a non-divergent 2D barotropic flow, stated as: "to optimally extract (kinetic) energy from the basic flow, a perturbation must be elongated locally along the axis of contraction of the basic deformation field" [physically interpreted by (Cai, 1992)]. In addition, (Cai and Mak, 1990) derived a whole set of energy equations for local perturbations in a zonally inhomogeneous baroclinic jet streak in a two-layer quasi-geostrophic beta-plane channel model [Cai and Mak, 1990, Eqs. (10-13)]. With these equations, (Mak, 1991) verified the "eddy straining" mechanism proposed by (Shutts, 1983), which argues that due to the impact of deformation, the transient eddy stretches along the north-south direction and compresses in the zonal direction, causing transformation of momentum and energy to the blocking flow and stimulating the blocking onset. Apart from discussing the perturbation growth mode within the deformation field, (Weiss, 1991) obtained a condition for the vorticity gradient increasing (Okubo-Weiss condition), to explain the vortex filament phenomenon with the conserved vertical vorticity equation in a barotropic and non-divergent atmosphere. According to the condition, the vertical vorticity gradient increases when the strength of the deformation square is larger than the vorticity square, and changes periodically when the vorticity square is larger than the deformation square. This theory has been used to explain the mechanism of the moat structure in tropical cyclones (Rozoff et al., 2006; Wang, 2008). These studies demonstrate well that deformation may play important roles in the physics of some weather phenomena, as emphasized by (Spensberger and Spengler, 2014), who stated that deformation may be used as a diagnostic tool in large-scale processes, e.g., low-level frontogenesis, evolution of high-level jets, orographic blocking, and Rossby wave breaking.

    While most of the above research is associated with the deformation of a barotropic atmosphere, some investigators have focused on the relation between deformation and disturbances in a baroclinic atmosphere. Bishop and Thorpe (1994a, 1994b) indicated that deformation has key effects on the unstable development of frontal waves, but may simultaneously suppress these waves. (Renfrew et al., 1997) also argued that the magnitude of stretching deformation of the environmental flow (Bishop, 1996a, 1996b) is crucial to the frontal stability. (Wang and Wu, 2001) found that a symmetric disturbance tends to develop under the ascending flow of the transverse circulation caused by deformational frontogenesis. With a 2D shallow-water model, (Jiang, 2011) investigated the formation and evolution of meso-β-scale low vortexes embedded in the deformational flow. In addition, (Jiang et al., 2011) proposed an interaction index between vortex and deformation flow, which was shown to be able to indicate the short-term direction of movement of a tropical cyclone. These investigations have provided significant insight into vortex development in a baroclinic atmosphere, implying an important role of deformation in baroclinic vortex evolution. However, most of these results were obtained from idealized numerical experiments, which are not easy to use in diagnostic analyses of the real atmosphere. For a barotropic atmosphere, both the non-divergent perturbation kinetic energy equation (Mak and Cai, 1989) and the Okubo-Weiss condition (Weiss, 1991) can be used to diagnose the possible roles of deformation in disturbance development. However, in a baroclinic atmosphere, a quantitative description and diagnosis of the impact of deformation on vortex development remains unrealized.

    Therefore, in this paper, a mathematical relation between deformation and the vertical vorticity tendency is built to describe and diagnose the impact of deformation on mesoscale low vortices in a baroclinic atmosphere. The corresponding formulation is presented in section 2. With this mathematical relation, a diagnostic study of the impact of deformation on vortex development in a heavy-rainfall case is reported in section 3. Finally, section 4 provides some concluding remarks.

2. Diagnostic method of vorticity development
  • In an adiabatic frictionless atmosphere, development of vertical vorticity is associated with the change of static stability, baroclinicity and horizontal vorticity under the restriction of conserved potential vorticity (PV). Since the commonly used vertical vorticity equation (referred to as the "classical vertical vorticity equation"; Wu, 2001; Holton, 2004) is derived from the momentum equation, these thermodynamic effects associated with static stability, baroclinicity and horizontal vorticity are thus not involved in the forcing of vertical vorticity development. In order to describe the impacts of these thermodynamic factors on the evolution of vertical vorticity and the associated inner dynamics, (Wu and Liu, 1999) derived a complete form of the vertical vorticity equation (CVE) from the PV equation and developed the slantwise vorticity development theory (SVD; Wu et al., 1995; Wu and Cai, 1997), which can explain the explosive increase of vertical vorticity during the frontal rainfall process. The PV equation and the corresponding CVE (Wu and Liu, 1999) in the z-coordinate can be written as \begin{equation} \dfrac{dP_{ E}}{dt}=\alpha{\nabla}\theta\cdot {F}+\alpha{\zeta}_{ a}\cdot{\nabla}Q , (1)\end{equation} and \begin{eqnarray} \dfrac{d}{dt}\zeta&=&-\beta v-\zeta_{ a}{\nabla}\cdot {v}-\dfrac{{\zeta}_{ h}\cdot{\nabla}_{ h}\theta}{\theta_{ z}} {\nabla}\cdot{v}-\dfrac{\zeta_{ a}}{\theta_{ z}}\dfrac{d}{dt}\theta_{ z}\nonumber\\[-0.5mm] &&-\dfrac{\zeta_{ h\theta}}{\theta_{ z}}\dfrac{d}{dt}|{\nabla}_{ h}\theta|-\dfrac{|{\nabla}_{ h}\theta|}{\theta_{ z}}\dfrac{d}{dt}\zeta_{ h\theta}\nonumber\\[-0.5mm] &&+\dfrac{1}{\theta_{ z}}({\nabla}\theta\cdot {F}+ {\zeta}_{ a}\cdot{\nabla}Q)(\theta_{ z}\ne 0) ,(2) \end{eqnarray} respectively, where P Eζ a·θ is the Ertel PV; α=1/ρ is the specific volume; ζ a=(∂w/∂y-∂v/∂z,∂u/∂z-∂w/∂x, ∂v/∂x-∂u/∂y+f) is the 3D absolute vorticity vector; θ=T(ps/p)R/cp is the potential temperature; T is temperature; p is pressure; ps is the reference pressure; R is the dry-air gas constant; cp is the specific heat at constant pressure; =(∂/∂x)i+(∂/∂y)j+ (∂/∂z)k is the three-dimensional gradient operator; F is the force of friction; Q is the diabatic heating, which contains latent heat release by water vapor condensation; ζ is the relative vertical vorticity; ζ a=ζ+f is the absolute vertical vorticity, where f is the Coriolis parameter; β=∂f/∂y is the longitudinal change of f; V=(u,v,w) is the 3D velocity vector; ζ h=(∂w/∂y-∂v/∂z,∂u/∂z- ∂w/∂x,0) is the horizontal vorticity vector; θ z=∂θ/∂z is the static stability; hθ=(∂/∂x)θ i+(∂/∂y)θ j is the horizontal potential temperature gradient; and \(\zeta_ h\theta=|\zeta_ h|\cos(\langle\zeta_ h,\nabla_ h\theta\rangle)\) is the projection of the horizontal vorticity vector ζ h along hθ.

    In Eq. (2), the first two terms on its right-hand side are the β effect term and the divergence term, which are also contained in classical vertical vorticity equation. The third to sixth terms are seen as thermodynamic processes contributing to vertical vorticity development, which are, respectively, the time change of specific volume, static stability, baroclinicity, and horizontal vorticity. In addition, friction dissipations and diabatic heating effects are contained in the seventh and eighth terms. In the previous almost 20 years since the development of SVD theory and the CVE, numerous studies (e.g., Cui et al., 2002; Meng et al., 2004; Li et al., 2005) have shown that static stability, vertical wind shear and baroclinicity all play important roles in the rapid (even explosive) increase of vertical vorticity under the slant of isentropic surfaces, which is usually associated with severe hazardous weather. For example, (Yu and Wu, 2001) showed that due to the steep orientation of moist isentropic surfaces, moist baroclinicity can trigger slantwise development of vertical vorticity, which may induce the rapid development of a tropical cyclone; (Wang et al., 2010) investigated a heavy snowfall case caused by the northward procession of a Bay of Bengal tropical storm, with MPV and SVD theory; and (Wu et al., 2013) and (Zheng et al., 2013) extended SVD theory into a generalization sense under a PV-θ view, and with this diagnostic tool they showed that horizontal vorticity and baroclinicity make a positive contribution to Tibetan Plateau low vortex development.

    In this paper, deformation strain is introduced into a diagnosis of vertical vorticity by the fifth term of Eq. (2); that is, the individual difference term of baroclinicity. This is based on the fact that deformation has been shown as the most important mechanism driving frontogenesis or the increase in baroclinicity, either by idealized numerical studies or by the frontogenesis function (e.g., Hoskins and Bretherton, 1972; Keyser and Anthes, 1982; Davies and Müller, 1988; Keyser et al., 1988). By coupling the CVE and frontogenesis function, a mathematical relation between deformation and vorticity can be built.

    The derivation begins with a generalization of the CVE into a moist baroclinic atmosphere due to the often close relation between strong vorticity and heavy rainfall; that is, the latent heat release caused by the water vapor phase change is considered. In the CVE, expressed by Eq. (2), latent heat is attributed to Q, which it is not possible to calculate accurately with conventional observations or reanalysis data. For the flexible use of general reanalysis data, the moist PV (MPV) in the p-coordinate is used as follows: \begin{equation} P_{ m}=-g({\nabla}\times{V}_{ h})\cdot{\nabla}\theta_{ e}=-g\left[{\zeta}_{ h}\cdot{\nabla}\theta_{ e} +(\zeta+f)\dfrac{\partial\theta_{ e}}{\partial p}\right] ,(3) \end{equation} where ζ h=(-∂v/∂p,∂u/∂p) is the horizontal vorticity vector, θ e=θexp(Lq/cpT) is the equivalent potential temperature, L is the latent heat, and q is the specific humidity. Note that the horizontal vorticity resulting from the vertical velocity is ignored in Eq. (3), mainly due to its small value compared with the horizontal vorticity resulting from vertical wind shear. However, this treatment is not applicable to a convective atmosphere where the magnitude of vertical velocity is comparable to the horizontal velocities. By the vorticity equation, the continuity equation and thermodynamic equation in the p-coordinate, the MPV equation can be derived (Wu et al., 1995) as \begin{equation} \dfrac{d}{dt}P_{ m}=-g{\nabla}\times {F}\cdot{\nabla}\theta_{ e}-g{\zeta}_{ a}\cdot{\nabla}Q_{ m} , (4)\end{equation} where g is gravitational acceleration, p=(∂/∂x)i+(∂/∂y)j is the gradient operator in the p-coordinate, and Q m is the diabatic heating rate apart from the latent heat. With Eq. (4), the complete form of the vertical vorticity equation in the p-coordinate can be derived: \begin{eqnarray} \dfrac{d}{dt}\zeta&=&-\beta v-\dfrac{\zeta_{ a}}{\theta_{ ep}}\dfrac{d}{dt}\theta_{ ep}-\dfrac{\zeta_{{ h}\theta_{ e}}}{\theta_{ ep}} \dfrac{d}{dt}|{\nabla}_{ p}\theta_{ e}|-\dfrac{|{\nabla}_{ p}\theta_{ e}|}{\theta_{ ep}}\dfrac{d}{dt}\zeta_{{ h}\theta_{ e}}\nonumber\\ &&+\dfrac{1}{\theta_{ ep}}{\nabla}\times {F}\cdot{\nabla}\theta_{ e}+\dfrac{1}{\theta_{ ep}}{\zeta}_{ a}\cdot{\nabla}Q_{ m} , (5)\end{eqnarray} where θ ep=∂θ e/∂p is convective stability, and ζ hθ e is the projection of the horizontal vorticity vector ζ h= (-∂v/∂p, ∂u/∂p) along pθ e. In Eq. (5), the first term on the right-hand side is the β effect term or the change of planetary vorticity, the second term is the temporal change term of the convective stability, the third term is the temporal change term of moist baroclinicity, the fourth term is the temporal change term of vertical wind shear, the fifth term is the friction term, and the sixth term is the effect of adiabatic processes.

    In a further step, the moist frontogenesis function is introduced, which is often used to diagnose the frontogenesis instead of the original frontogenesis function when there is precipitation (Bluestein, 1993): \begin{eqnarray} F&\!=\!&\dfrac{d}{dt}|{\nabla}_{ p}\theta_{ e}|=F_{ G1}+F_{ G2}+F_{ G3}+F_{ G4} ;(6)\\[0.5mm] F_{ G1}&\!=\!&-\dfrac{D}{2}|{\nabla}_{ p}\theta_{ e}| ;(7)\\[0.5mm] F_{ G2}&\!=\!&-\!\dfrac{1}{2|{\nabla}_{ p}\theta_{ e}|}\!\left[E_{ st}\!\left(\dfrac{\partial\theta_{ e}}{\partial x}\right)^2 \!\!+\!2E_{ sh}\dfrac{\partial\theta_{ e}}{\partial x}\dfrac{\partial\theta_{ e}}{\partial y} \!-\!E_{ st}\!\left(\dfrac{\partial\theta_{ e}}{\partial y}\right)^2\!\right]\nonumber\\[0.5mm] &\!=\!&\dfrac{|{\nabla}_{ p}\theta_{ e}|}{2}|E|\cos 2\gamma ;(8)\\[0.5mm] F_{ G3}&\!=\!&-\dfrac{1}{|{\nabla}_{ p}\theta_{ e}|}\left(\dfrac{\partial\omega}{\partial x}\dfrac{\partial\theta_{ e}}{\partial x} +\dfrac{\partial\omega}{\partial y}\dfrac{\partial\theta_{ e}}{\partial y}\right)\dfrac{\partial\theta_{ e}}{\partial p} ;(9)\\[0.5mm] F_{ G4}&\!=\!&\dfrac{1}{|{\nabla}_{ p}\theta_{ e}|}\left[\dfrac{\partial\theta_{ e}}{\partial x} \dfrac{\partial Q_{ m}}{\partial x}+\dfrac{\partial\theta_{ e}}{\partial y}\dfrac{\partial Q_{ m}}{\partial y}\right] . (10)\end{eqnarray} Here, F G1 is the divergence term, F G2 is the deformation term, F G3 is the tilting term, and F G4 is the diabatic term. D is horizontal divergence, E st is horizontal stretching deformation, E sh is horizontal shearing deformation, and γ is the angle between the equivalent potential temperature contours and the dilatation axis, which controls the effect of deformation on front development. When γ<45°, there is deformational moist frontogenesis, and when γ>45° there is deformational moist frontolysis.

    Equation (6) is directly substituted into the moist baroclinicity term in the CVE [Eq. (5)], and the CVE in p-coordinates can thus be written as \begin{equation} \dfrac{\partial}{\partial t}\zeta =F_{\zeta 1}+F_{\zeta 2}+F_{\zeta 3}+F_{\zeta 4}+F_{\zeta 5} , (11)\end{equation} where the β effect, frictions and diabatic heating have been ignored, and:

    (1) Fζ 1=-V h$\nabla$pζ is horizontal advections of vertical vorticity;

    (2) Fζ 2=-ω(∂ζ/∂p) is vertical advection of vertical vorticity;

    (3) Fζ 3=-(ζ a ep)(dθ ep/dt) is the time variation term of convective stability;

    (4) Fζ 4=-(ζ hθ e ep)(d|∇ pθ e|/dt)=-(ζ hθ e ep) (F G1+F G2+F G3)=Fζ 41+Fζ 42+Fζ 43 is the time variation term of moist baroclinicity, in which \begin{equation} F_{\zeta 41}=-\dfrac{\zeta_{{ h}\theta_{ e}}}{\theta_{ ep}}F_{ G1} =\dfrac{\zeta_{ h}\cdot\nabla_{ p}\theta_{ e}}{2\theta_{ ep}}D , (12)\end{equation} is the vertical vorticity development caused by divergence-induced moist baroclinicity variation, \begin{equation} F_{\zeta 42}=-\dfrac{\zeta_{{ h}\theta_{ e}}}{\theta_{ ep}}F_{ G2}=-\dfrac{\zeta_{ h}\cdot\nabla_{ p}\theta_{ e}} {2\theta_{ ep}}|E|\cos 2\gamma , (13)\end{equation} is the impact of deformation-induced moist baroclinicity variation on vertical vorticity development, and \begin{equation} F_{\zeta 43}=-\dfrac{\zeta_{{ h}\theta_{ e}}}{\theta_{ ep}}F_{ G3}=\dfrac{\zeta_{{ h}\theta_{ e}}}{\theta_{ ep}} \dfrac{1}{|\nabla_{ p}\theta_{ e}|}\left(\dfrac{\partial\omega}{\partial x} \dfrac{\partial\theta_{ e}}{\partial x}+\dfrac{\partial\omega}{\partial y}\dfrac{\partial\theta_{ e}}{\partial y}\right) \dfrac{\partial\theta_{ e}}{\partial p} , (14)\end{equation} is the vertical vorticity development caused by the horizontal inhomogeneous distribution of vertical motions;

    (5) Fζ 5=-(|∇ pθ e|/θ ep)(d/dt)ζ hθ e is the time variation term of vertical wind shear (or horizontal vorticity).

    In all of the above forcing terms in the CVE, it can be seen that deformation is directly contained in the baroclinicity variation-related term Fζ42, which then affects development of vertical vorticity. From Fζ42, it is shown that whether deformation can induce vorticity development not only depends on deformation itself, but also the atmospheric thermodynamic background; that is, the convective stability, vertical wind shear and baroclinicity. These are explicitly diagnosed in the following case study.

3. Case study
  • With the above theoretical theorems [Eqs. (13) and (11)], the impact of deformation on vorticity development is studied in a heavy-rainfall event that occurred in a strong moist baroclinic environment. The data for the analysis are from the NCEP (National Centers for Environmental Prediction) global reanalysis dataset, with a horizontal resolution of 0.5°× 0.5°. Figure 1 shows the precipitation distribution of the studied case, which occurred during 20-22 July 2012. As shown in Fig. 1, the precipitation has an evident belt structure, oriented in a northeast-southwest direction. On 20 July (Fig. 1a), the rainbelt is mainly located northwest of China, with three heavy-precipitation centers in Inner Mongolia (labelled A), the joint areas ofShaanxi, Ningxia and Gansu provinces (B), and north of the Sichuan Basin (C). On 21 July (Fig. 1b), the rainbelt displaces southeastward and elongates, also with three heavy-precipitation centers. In Fig. 1b, center A, which has the heaviest precipitation, covers Beijing, Tianjin and part of Hebei. On this day, the averaged daily precipitation amount in Beijing reaches up to 170 mm, and the total precipitation amount reaches up to 460 mm in Hebei town of Fangshan district, which induced great economic losses and casualties. On 22 July, with the eastward movement of the rainbelt, precipitation decreases and the belt structure mainly persists in East and Northeast China.

    Figure 2 shows the distributions of some basic variables in typical levels at 1200 UTC 21 July 2012, illustrating the weather configuration of this case. Corresponding to the orientation and location of the 6-h accumulated precipitation belt at 1200 UTC, there is a northeast-southwest oriented jet stream (shaded areas of Fig. 2a; wind speed >30 m s-1) in the 200 hPa upper level and a moist strong baroclinic zone (cold front; black lines in Fig. 2a) at 700 hPa in the lower troposphere. The heavy rainfall is closely related to this cold front. The front leads the cold air mass southward, and encounters warm-moist air flow from low latitudes, which induces strong precipitation along the cold front. In the mid-level (Fig. 2b), the precipitation area is characterized by deformational flow, with cyclonic wind shear (low trough at the bottom of the cyclone, centered over Lake Baikal) on the north side of the rainbelt, which causes cold-air intrusion in the mid-level and tends to strengthen the rainfall. In addition, the 500 hPa southwesterly prevailing over the areas of Hebei, Beijing and Tianjin also leads meso-γ scale convective cells moving northeasterly. This so-called steering flow has an impact on the `echo-training' mechanism of the `Beijing 721 heavy rainfall' event. As apparent in Fig. 2, another evident feature of the weather pattern is a meso-β scale low vortex (about 500 km in length) in the most severe precipitation area (about 41°N, 116.5°E). The low vortex is collocated with the low-level jet at 700 hPa (wind speed >12 m s-1; blue lines in Fig. 2a). In Fig. 2a, the low-level geopotential height (red lines) shows the mesoscale low vortex to be connected to the large vortex in the vicinity of Lake Baikal, overlaying the strong baroclinic zone or cold front. Strong coherent rotation caused by the low vortex plays a significant role in the persistence of the precipitating system, which provides a favorable environment for the concentration of energy and transport. In the whole process of the studied rainfall case, the strongest precipitation center is collocated with, and moves with, the vortex. As shown in Fig. 3, which gives the movement of the low vortex with the three-day averaged geopotential height at 700 hPa as a background, the low vortex develops and moves in a typical saddle-shaped field with two highs and two lows alternatively distributed. In the high latitudes, a high ridge is located west of Lake Baikal, while a low vortex is located west of the lake. In the low latitudes is a low vortex over the Bay of Bengal, south of the Tibetan Plateau (approximately 80°-90°E), and a subtropical high over the east of China over the ocean. A large-scale deformation field then forms over the north and northwest of China, plus the blocking effect of the Tibetan Plateau. As the low vortex initializes on the northeast edge of the Tibetan Plateau, it moves northeastward along the dilation axis of the large-scale deformation field, which influences the provinces of Shaanxi, Shanxi, Hebei, Liaoning and Jilin, successively. The saddle-shaped weather pattern is a typical deformation field in the atmosphere. Evolution of the low vortex in the saddle-shaped weather pattern, or in the large-scale deformation field, shows that deformation may have played an important role in the vortex's development, which has been noted in several other studies. For example, with an idealized numerical experiment, (Jiang, 2011) showed that the deformational field has an important impact on the formation of a meso-β scale vortex. Through statistical analysis of the boundary deformation fields during the Mei-yu period, (Deng, 1986) stated that deformation can be applied to indicate the movement and development of low vortex systems. However, to explicitly identify the relation of the deformation field and vortex development, further quantitative diagnosis would be necessary. In the following part of the study, the impact of deformation on the evolution of the low vortex in the strong baroclinic frontal zone is diagnosed.

    Figure 1.  The 24-h observational accumulated precipitation during 20-22 July 2012 (units: mm).

    Figure 2.  Horizontal wind speed at 200 hPa (colored areas; units: m s$^-1$), equivalent potential temperature (black solid lines; units: K), geopotential height (red solid lines; units: 10 gpm) and horizontal wind speed ($>$12 m s$^-1$; blue solid lines; units: m s$^-1$) at 700 hPa. (b) Horizontal wind field (black arrows) and geopotential height (red solid lines) at 500 hPa, and 6-h observational accumulated precipitation (colored areas; units: mm) at 1200 UTC 21 July 2012.

    Figure 3.  Three-day time-averaged geopotential height at 700 hPa during 20-22 July 2012 (units: 10 gpm). The black dot is the location of the low vortex center at different times, and the lines between indicate the direction of movement of the low vortex.

  • According to the interactions of the low vortex and the cold front, the lifecycle of the vortex is decomposed into four stages: the initial stage, development stage, maturation stage, and dissipating stage (Fig. 4 and Table 1). In the initial stage at 1200 UTC 20 July (Fig. 4a), the strong baroclinic frontal zone is on the north border of China (about 42°N). The low vortex, denoted by the large positive vertical vorticity area, forms in the warm sector of the front, influenced by the prefrontal warm moist flow. After initiation, with the southward intrusion of the cold front, the vortex is gradually superposed over the cold front and develops substantially (Fig. 4b). Then, the vortex (large positive vorticity anomalies) moves northeastward within the front and enters the Beijing area at 1200 UTC 21 July (Fig. 4c), when severe heavy precipitation occurs over Beijing. Meanwhile, the vortex reaches its maturation stage. Subsequently, with longitudinal enlargement of the cold front, the vortex gradually separates with the front and dissipates (Fig. 4d), showing a tendency to merge with the high-latitude cyclone (figure omitted).

    Figure 4.  Vortex evolution denoted by vertical vorticity (colored areas; units: 10$^-5$ s$^-1$): (a) initial stage at 1200 UTC 20 July; (b) development stage at 0000 UTC 21 July; (c) mature stage at 1200 UTC 21 July; (d) dissipating stage at 0600 UTC 22 July 2012. The black solid lines are equivalent potential temperature.

  • The deformation related term Fζ42 [Eq. (13)] is calculated over every stage of the vortex to diagnose the impact of deformation on the vortex's development. Figure 5 shows the vertical distributions of Fζ42 along the vortex center. As shown in Fig. 5a, in the initial stage, the positive vorticity area over the formation location (37.5°N, 105.5°E) of the vortex stretches from the surface (approximately 800 hPa over the terrain) up to 600 hPa, with the center close to the surface (about 16× 10-5 s-1). The deformation-related term Fζ42 mainly presents negative values over the positive vorticity area, which implies that deformation suppresses the increase of positive vorticity and thus acts against the vortex's initiation. As shown in Fig. 5b, in the development stage of the vortex at 0000 UTC 21 July, positive vertical vorticity presents a significant increase and forms a coherent vorticity column, stretching from the surface up to 500 hPa, with the center at about 750 hPa (30× 10-5 s-1). In this stage, with the southward intrusion of the cold front into the vortex, the impact of deformation on vertical vorticity development changes evidently. In Fig. 5b, over the vorticity column, the deformation-related term Fζ42 shows positive values above 750 hPa and negative values below 750 hPa. The positive center (60× 10-9 s-2) is approximately at the 700 hPa level, with the intensity much stronger than the negative center. This means that, in the development stage, deformation changes to increase the vertical vorticity and promote the vortex's development, with the most evident effect at 700 hPa. After the explosive genesis in the development stage, the vortex evolves into a slow-developing maturation stage, as shown in Fig. 5c. In this stage, the vortex enters North China and influences the Heibei-Beijing-Tianjin area, inducing extreme precipitation. At 1200 UTC 21 July, the vortex center is at about 41°N and the associated vorticity column shows a downward sliding to the plain, with the length and height much larger than in the previous two stages. Over the vorticity column, the deformation-related term Fζ42 shows all positive values, stretching from the boundary (900 hPa) up to 600 hPa. This means, in the maturation stage, deformation also plays a role in promoting the vortex's development, albeit with a much weaker intensity. After the maturation stage, the vortex gradually moves out of North China and enters the Northeast Plain. Along the vortex center at about 43°N, the positive vorticity column persists, with the deformation-related term Fζ42 scattered in its distribution, but shows an increasing-vorticity pattern north of 44°N (black solid curve in Fig. 5d).

    Figure 5.  Vertical distribution of the deformation-related vorticity development term $F_\zeta 42$ (units: 10$^-9$ s$^-2$) along the longitude passing through the vortex center (Table 1): (a) initial stage at 1200 UTC 20 July; (b) development stage at 0000 UTC 21 July; (c) mature stage at 1200 UTC 21 July; (d) dissipating stage at 0600 UTC22 July 2012. The black solid lines are negative values; red lines are positive values. The colored areas are vertical vorticity. (units: 10$^-5$s$^-1$)

    Figure 6.  Vertical vorticity budget in the vortex center area obtained by Eq. (11) during 20-22 July 2012: (a) sketch map of every term; (b) averaged budget over (37.0°-38.0°N, 105.0°-106.0°E) at 700 hPa at 1200 UTC 20 July; (c) averaged budget over (38.0°-39.0°N, 108.5°-109.5°E) at 700 hPa at 0000 UTC 21 July; (d) averaged budget over (40.5°-41.5°N, 116.0°-117.0°E) at 800 hPa at 1200 UTC 21 July; (e) averaged budget over (43.5°-44.5°N, 123.5°-124.5°E) at 850 hPa at 1200 UTC 22 July. The red (blue) box is the main term increasing (decreasing) positive vertical vorticity. The blue dotted box is the deformation-related term.

    Figure 7.  Total deformation (left column; red lines; units: 10$^-5$ s$^-1$) and deformation tick marks (right column; black short line; units: 10$^-5$ s$^-1$). The shaded areas are vertical vorticity; the red lines in the right column are equivalent potential temperature (units: 1 K); the black box is the critical area of the vortex; and the black dots indicate the vortex center.

    Figure 8.  Vertical distribution of $|E|\cos 2\gamma$ (red lines; units: 10$^-5$ s$^-1$) along the vortex center: (a) 105.5°E at 1200 UTC 20 July; (b) 109°E at 0000 UTC 21 July; (c) 116.5°E at 1200 UTC 21 July; and (d) 124°E at 1200 UTC 22 July. The black solid lines are $\theta_ e$ (units: K) contours, and the shaded areas are vertical vorticity (units: 10$^-5$ s$^-1$). The grey bars are 6-h accumulated precipitation.

    Figure 9.  Vertical distribution of $P_ m2$ (black lines; units: 10$^-1$ PVU) along the vortex center: (a) 105.5°E at 1200 UTC 20 July; (b) 109°E at 0000 UTC 21 July; (c) 116.5°E at 1200 UTC 21 July; and (d) 124°E at 1200 UTC 22 July. The gray solid lines are $\theta_ e$ (units: K) contours, and the shaded areas are zonal wind (units: m s$^-1$). The ellipses are the locations of the vorticity columns.

    The above analysis highlights the fact that, in the real atmosphere, deformation can indeed induce the development of vertical vorticity, although the specific effects may depend on the temporal dynamic and thermodynamic configurations. The relative importance of the deformation-related term compared with other terms in the CVE is estimated by calculating the budget of every term in the CVE over the vortex center area in its each stage. Note that the levels that are chosen to give the calculation are a little different in different stages due to the vertical variation of the deformation-related term with time (Fig. 5). In Fig. 6, it is shown that the primary factor that contributes to the vortex's development presents a significant difference. In the initial stage (Fig. 6b), the total forcing [sum of the terms on the right-hand side of Eq. (11)] is positive, which corresponds to the increase in vertical vorticity and formation of the vortex. The vertical wind shear term Fζ 5 makes the largest contribution to the vortex's formation, while the convective stability-variation term Fζ 3 suppresses the vortex's formation. The moist baroclinicity-variation term makes almost no contribution to the vortex's formation, although it has positive values and tends to increase the vertical vorticity. The deformation-related term Fζ42 displays suppression of the vortex's development in this stage, but its magnitude is quite small. In the development stage (Fig. 6c), the vertical wind shear term Fζ 5 and the convective stability-variation term Fζ 3 are still the two leading terms that influence the vortex's development, but with the suppression effect by Fζ 3 becoming almost twofold greater than the promotion effect by Fζ 5. However, the total forcing is still positive and the low vortex develops. This is mostly due to the increased contribution from the moist baroclinicity-variation term, which is mainly caused by the increase of the deformation-related term (Figs. 6b-c). In the maturation stage, positive total forcing maintains in the vortex center, to keep the vertical vorticity increasing. The vertical wind shear term Fζ5 is still the main factor driving the vortex's development, while the vertical advection term Fζ 2 becomes the main suppression term instead of Fζ 3. At this time, the deformation-related term still promotes the vortex's development, but with a smaller contribution compared to the last stage. In the dissipating stage (Fig. 6e), the total forcing of vertical vorticity in the vortex center becomes negative. On the contrary to the maturation stage, the vertical wind shear term turns into a negative factor for vertical vorticity increasing, while the vertical advection term begins to maintain the vortex; that is, resisting the dissipation of the vortex. The deformation-related term also shows a promotion of increased vertical vorticity.

    From the above analysis, the impact of deformation on vortex evolution and its magnitude is quite different in the various vortex stages. Both the vertical distributions of the deformation-related term and the vorticity budget show that, during the whole lifecycle of the low vortex, deformation has the most important effect on increasing vertical vorticity during the vortex development stage; that is, when the cold front intrudes southward into the vortex. In order to explain the specific dynamic and thermodynamic structures that are favorable to the deformation in promoting the vortex's development, the deformation-related term is the focus of discussion in the next subsection.

  • The deformation related term [Eq. (13)] is rewritten as follows for further discussion: \begin{eqnarray} F_{\zeta 42}&=&\dfrac{P_{ m2}|E|\cos 2\gamma}{2\theta_{ ep}} ,(15)\\ P_{ m2}&=&-{\zeta}_{ h}\cdot\nabla_{ p}\theta_{ e}=\dfrac{\partial v}{\partial p} \dfrac{\partial\theta_{ e}}{\partial x}-\dfrac{\partial u}{\partial p}\dfrac{\partial \theta_{ e}}{\partial y} , (16)\end{eqnarray} where the baroclinic MPV (P m2; Wang et al., 1996) is applied to substitute the coupling of the vertical wind shear and the moist baroclinicity. Thus, from Eq. (15), the factors that influence the effect of deformation on vorticity include convective stability θ ep, the baroclinic MPV (P m2) and the deformation |E|cos 2γ (γ is the angle of the dilation axis of the deformation field and the θ e contours). These factors are together analyzed to determine their effect on the vortex's development. Figure 7 gives the horizontal distribution of the total deformation |E| and the deformation tick marks, which represent the magnitude and direction of the dilation axis of the deformation field. According to Figs. 7a, c, e and g, during the whole lifecycle of the vortex, strong deformation areas are always superposed over the vorticity belts. The magnitude of the total deformation in the vortex center in the different stages are ∼6× 10-5 s-1, ∼15× 10-5 s-1, ∼18× 10-5 s-1 and ∼ 15× 10-5 s-1, which implies that, after initiation, the total deformation of the central vortex area maintains a steady state. Since the total deformation is always positive, the sign of |E|cos 2γ thus depends on the angle between the deformation dilation axis and the θ e contours, which can be seen from Figs. 7b, d, f and h. Because of the southwest warm, moist flow into the inner land, the vortex is basically under the control of a warm ridge, seen from the convex θ e contours. In the warm ridge, the deformation tick marks are basically aligned with the θ e contours. West of the ridge, the tick marks are southwest-northeast or west-east oriented, while east of the ridge the tick marks become northwest-southeast oriented with the curve of the θ e contours. This means, in the large vorticity areas, deformation always presents frontogenesis; that is, |E|cos 2γ>0, which can also be seen from Fig. 8. In the vertical cross sections along the vortex center (Fig. 8), the deformation term |E|cos 2γ>0 is always positive during the whole lifecycle of the vortex, either when the front intrudes into the vortex or not. Figure 9 gives the vertical distributions of the baroclinic MPV (P m2) along the vortex center. From Fig. 9, P m2 mainly shows negative values over the vortex during its whole lifecycle. This is mostly due to the superposition of the high-level westerly over the low-level easterly (∂u/∂p<0) and the moist baroclinicity caused by the encountering of high-latitude cold air mass and low-latitude warm moist flow (∂θ e/∂y<0). The configuration of the deformation term and the baroclinic MPV make the numerator of Eq. (15) mainly present negative values over the vortex area, which implies that whether or not the deformation increases the vertical vorticity depends on convective stability. In the initial stage, the vortex lies in the prefrontal unstable stratification (Fig. 10a; θ ep>0); thus, the deformation-related term suppresses the vortex initiation (Fig. 5a). In the development stage, the front gradually intrudes into the vortex. The atmosphere is stable above 750 hPa and unstable below 750 hPa, which means a neutral stratification exists between these two air layers. This results in the deformation-related term showing a large positive-value center in the near-neutral layer, deriving a vertical vorticity increase above the neutral layer and decreasing it below the layer (Fig. 5b). This neutral stratification is also the main reason that deformation plays a most evident role in the vorticity increase during the vortex development stage, compared with other stages. In the maturation stage, the front and the vortex are superposed, and the atmosphere in the vortex is completely stable (θ ep<0), making the deformation-related term increase the vertical vorticity but with a smaller magnitude.

    From the above analysis, it can be concluded that the real effect of deformation on the vertical vorticity trend in fact takes place through the deformational frontogenesis process, but under the restriction of the baroclinic PV and the atmospheric stability. The inner physical process can be qualitatively analyzed from the MPV. Assuming the stability of the atmosphere (θ ep) does not change during the deformational frontogenesis process, the increase of |∇θ e| will thus make the inclination of the moist isentropic surfaces increase (|∇θ e ep|). According to the expression of the vertical vorticity that is obtained from MPV [Eq. (3)], \begin{equation} \zeta=-\dfrac{P_{ m}}{g\frac{\partial\theta_{ e}}{\partial p}}+\dfrac{{\zeta}_{ h}\cdot{\nabla}\theta_{ e}} {\frac{\partial\theta_{ e}}{\partial p}}-f=-\dfrac{P_{ m}}{g\frac{\partial\theta_{ e}}{\partial p}}+ |\theta_{ ep}|\dfrac{\zeta_{{ h}\theta_{ e}}}{\theta_{ ep}}\left|\dfrac{{\nabla}\theta_{ e}}{\theta_{ ep}}\right|-f , (17)\end{equation} the increase of the inclination of the isentropic surfaces will then induce the vertical vorticity development, as long as the atmosphere satisfies certain conditions (ζ_ hθ eθ ep>0). In a certain sense, this is also part of the SVD theory, except that the factor (mainly deformation) that drives the isentropic surface inclination is considered.

    Figure 10.  Vertical distribution of $\theta_ ep$ (red lines; units: 10$^-1$ K hPa$^-1$) along the vortex center: (a) 105.5°E at 1200 UTC 20 July; (b) 109°E at 0000 UTC 21 July; (c) 116.5°E at 1200 UTC 21 July. The black solid lines are $\theta_ e$ (units: K) contours, and the shaded areas are vertical vorticity (units: 10$^-5$ s$^-1$).

4. Conclusion
  • A mathematical relation between deformation and vertical vorticity development is built by using the frontogenesis function as a substitute for the baroclinic variation term in the CVE, which is derived from the MPV. The relation provides valid evidence and diagnostics for the impact of deformation on mesoscale vortex development. Through the deformation-related term in the CVE, it is shown that the impact of deformation on increasing vertical vorticity not only relates to the deformation itself, but also needs the proper configuration of convective stability, moist baroclinicity and vertical wind shear. With this term, the impact of deformation on the development of a mesoscale vortex in a frontal heavy-rainfall case, which occurred during 20-22 July 2012, is diagnosed. The elements in this deformation-related term, including the deformation-induced baroclinic variation (|E|cos 2γ), the convective stability (θ ep) and the baroclinic part of MPV (P m2), are diagnosed. The results show that, because the baroclinic MPV usually shows negative values in the vortex due to the influence of the front, and deformation often makes a positive contribution to the moist baroclinicity increase (|E|cos 2γ>0), whether or not deformation promotes the vortex's development mostly depends on the stability of the baroclinic moist atmosphere. In the initial stage of the vortex, the atmosphere is convectively unstable, and deformation suppresses vortex development. In the development stage, as the cold front intrudes into the vortex, the stratification of the atmosphere in the vortex becomes stable above 750 hPa and remains unstable below 750 hPa, which causes the deformation to promote increasing vertical vorticity above 750 hPa and, thus, the vortex's development. In the maturation and dissipating stages, the stratifications in the vortex are all stable, and deformation promotes the vortex's development from its bottom to the top. By comparing the magnitudes of different terms in the CVE during the whole process of the vortex's evolution, it is in the vortex development stage that deformation shows a most significant impact on increasing vorticity. This is mostly due to a near-neutral level between a stable atmosphere and an unstable atmosphere, which makes the deformation-related term in the CVE much larger in the vortex development stage than in other stages.

Reference

Catalog

    /

    DownLoad:  Full-Size Img  PowerPoint
    Return
    Return