Journal articles on the topic 'Baroclinic'

To see the other types of publications on this topic, follow the link: Baroclinic.

Create a spot-on reference in APA, MLA, Chicago, Harvard, and other styles

Select a source type:

Consult the top 50 journal articles for your research on the topic 'Baroclinic.'

Next to every source in the list of references, there is an 'Add to bibliography' button. Press on it, and we will generate automatically the bibliographic reference to the chosen work in the citation style you need: APA, MLA, Harvard, Chicago, Vancouver, etc.

You can also download the full text of the academic publication as pdf and read online its abstract whenever available in the metadata.

Browse journal articles on a wide variety of disciplines and organise your bibliography correctly.

1

Donohoe, Aaron, and David S. Battisti. "Causes of Reduced North Atlantic Storm Activity in a CAM3 Simulation of the Last Glacial Maximum." Journal of Climate 22, no. 18 (September 15, 2009): 4793–808. http://dx.doi.org/10.1175/2009jcli2776.1.

Full text
Abstract:
Abstract The aim of this paper is to determine how an atmosphere with enhanced mean-state baroclinity can support weaker baroclinic wave activity than an atmosphere with weak mean-state baroclinity. As a case study, a Last Glacial Maximum (LGM) model simulation previously documented to have reduced baroclinic storm activity, relative to the modern-day climate (simulated by the same model), despite having an enhanced midlatitude temperature gradient, is considered. Several candidate mechanisms are evaluated to explain this apparent paradox. A linear stability analysis is first performed on the jet in the modern-day and the LGM simulation; the latter has relatively strong barotropic velocity shear. It was found that the LGM mean state is more unstable to baroclinic disturbances than the modern-day mean state, although the three-dimensional jet structure does stabilize the LGM jet relative to the Eady growth rate. Next, feature tracking was used to assess the storm track seeding and temporal growth of disturbances. It was found that the reduction in LGM eddy activity, relative to the modern-day eddy activity, is due to the smaller magnitude of the upper-level storms entering the North Atlantic domain in the LGM. Although the LGM storms do grow more rapidly in the North Atlantic than their modern-day counterparts, the storminess in the LGM is reduced because storms seeding the region of enhanced baroclinity are weaker.
APA, Harvard, Vancouver, ISO, and other styles
2

Pierrehumbert, R. T., and K. L. Swanson. "Baroclinic Instability." Annual Review of Fluid Mechanics 27, no. 1 (January 1995): 419–67. http://dx.doi.org/10.1146/annurev.fl.27.010195.002223.

Full text
APA, Harvard, Vancouver, ISO, and other styles
3

Kantha, Lakshmi H., and Craig C. Tierney. "Global baroclinic tides." Progress in Oceanography 40, no. 1-4 (January 1997): 163–78. http://dx.doi.org/10.1016/s0079-6611(97)00028-1.

Full text
APA, Harvard, Vancouver, ISO, and other styles
4

REZNIK, GREGORY M., and GEORGI G. SUTYRIN. "Baroclinic topographic modons." Journal of Fluid Mechanics 437 (June 22, 2001): 121–42. http://dx.doi.org/10.1017/s0022112001004062.

Full text
Abstract:
The theory of solitary topographic Rossby waves (modons) in a uniformly rotating two-layer ocean over a constant slope is developed. The modon is described by an exact, form-preserving, uniformly translating, horizontally localized, nonlinear solution to the inviscid quasi-geostrophic equations. Baroclinic topographic modons are found to translate steadily along contours of constant depth in both directions: either with negative speed (within the range of the phase velocities of linear topographic waves) or with positive speed (outside the range of the phase velocities of linear topographic waves). The lack of resonant wave radiation in the first case is due to the orthogonality of the flow field in the modon exterior to the linear topographic wave field propagating with the modon translation speed, that is impossible for barotropic modons. Another important property of a baroclinic topographic modon is that its integral angular momentum must be zero only in the bottom layer; the total angular momentum can be non-zero unlike for the beta-plane modons over flat bottom. This feature allows modon solutions superimposed by intense monopolar vortices in the surface layer to exist. Explicit analytical solutions for the baroclinic topographic modons with piecewise linear dependence of the potential vorticity on the streamfunction are presented and analysed.
APA, Harvard, Vancouver, ISO, and other styles
5

Cehelsky, Priscilla, and Ka Kit Tung. "Nonlinear Baroclinic Adjustment." Journal of the Atmospheric Sciences 48, no. 17 (September 1991): 1930–47. http://dx.doi.org/10.1175/1520-0469(1991)048<1930:nba>2.0.co;2.

Full text
APA, Harvard, Vancouver, ISO, and other styles
6

Fleming, Rex J. "Explosive Baroclinic Instability." Journal of the Atmospheric Sciences 71, no. 6 (May 30, 2014): 2155–68. http://dx.doi.org/10.1175/jas-d-13-0323.1.

Full text
Abstract:
Abstract A low-order general circulation model contains all the elements of baroclinic instability, including differential heating to drive the mean zonal shear flow against dissipation. Simulations exhibit vacillation ending in fixed-point solutions and chaotic solutions with significant amplifications of the baroclinic cycles compared to those of vacillation. The chaos sensitivity to initial conditions, covering a broad landscape of initial values, demands analysis of why the chaos occurs and its impact on subsequent storm intensity. Three attractors found in the dynamic system are important. One attractor is the stable fixed-point solution—the ultimate destination of a vacillation trajectory. A second attractor represents an unstable zonal solution. Though this dynamic system is bound, some trajectories get extremely close to the unstable, but strongly attracting, zonal solution. It is while traversing such a trajectory that the buildup of available potential energy is such to allow subsequent explosive baroclinic instability to develop. The roots of the characteristic matrix of the dynamic system are examined at every time step. A single critical value of one of the roots is found to be the cause of the chaos for a given value of the differential heating H. The system becomes more stable with increased values of H; vacillation is stronger and more prominent, and the critical value for chaos increases with H. When chaos does occur, it is stronger and more explosive.
APA, Harvard, Vancouver, ISO, and other styles
7

Zhang, Zheen, Xueen Chen, and Thomas Pohlmann. "The Impact of Fortnightly Stratification Variability on the Generation of Baroclinic Tides in the Luzon Strait." Journal of Marine Science and Engineering 9, no. 7 (June 26, 2021): 703. http://dx.doi.org/10.3390/jmse9070703.

Full text
Abstract:
The impact of fortnightly stratification variability induced by tide–topography interaction on the generation of baroclinic tides in the Luzon Strait is numerically investigated using the MIT general circulation model. The simulation shows that advection of buoyancy by baroclinic flows results in daily oscillations and a fortnightly variability in the stratification at the main generation site of internal tides. As the stratification for the whole Luzon Strait is periodically redistributed by these flows, the energy analysis indicates that the fortnightly stratification variability can significantly affect the energy transfer between barotropic and baroclinic tides. Due to this effect on stratification variability by the baroclinic flows, the phases of baroclinic potential energy variability do not match the phase of barotropic forcing in the fortnight time scale. This phenomenon leads to the fact that the maximum baroclinic tides may not be generated during the maximum barotropic forcing. Therefore, a significant impact of stratification variability on the generation of baroclinic tides is demonstrated by our modeling study, which suggests a lead–lag relation between barotropic tidal forcing and maximum baroclinic response in the Luzon Strait within the fortnightly tidal cycle.
APA, Harvard, Vancouver, ISO, and other styles
8

Ji, Xuan, J. David Neelin, and C. Roberto Mechoso. "Baroclinic-to-Barotropic Pathway in El Niño–Southern Oscillation Teleconnections from the Viewpoint of a Barotropic Rossby Wave Source." Journal of the Atmospheric Sciences 73, no. 12 (November 29, 2016): 4989–5002. http://dx.doi.org/10.1175/jas-d-16-0053.1.

Full text
Abstract:
Abstract The baroclinic-to-barotropic pathway in ENSO teleconnections is examined from the viewpoint of a barotropic Rossby wave source that results from decomposition into barotropic and baroclinic components. Diagnoses using the NCEP–NCAR reanalysis are supplemented by analysis of the response of a tropical atmospheric model of intermediate complexity to the NCEP–NCAR barotropic Rossby wave source. Among the three barotropic Rossby wave source contributions (shear advection, vertical advection, and surface drag), the leading contribution is from shear advection and, more specifically, the mean baroclinic zonal wind advecting the anomalous baroclinic zonal wind. Vertical advection is the smallest term, while surface drag tends to cancel and reinforce the shear advection in different regions through damping on the baroclinic mode, which spins up a barotropic response. There are also nontrivial impacts of transients in the barotropic wind response to ENSO. Both tropical and subtropical baroclinic vorticity advection contribute to the barotropic component of the Pacific subtropical jet near the coast of North America, where the resulting barotropic wind contribution approximately doubles the zonal jet anomaly at upper levels, relative to the baroclinic anomalies alone. In this view, the barotropic Rossby wave source in the subtropics simply arises from the basic-state baroclinic flow acting on the well-known baroclinic ENSO flow pattern that spreads from the deep tropics into the subtropics over a scale of equatorial radius of deformation. This is inseparably connected to the leading deep tropical Rossby wave source that arises from eastern Pacific climatological baroclinic winds advecting the tropical portion of the same ENSO flow pattern.
APA, Harvard, Vancouver, ISO, and other styles
9

Canals, Miguel, Geno Pawlak, and Parker MacCready. "Tilted Baroclinic Tidal Vortices." Journal of Physical Oceanography 39, no. 2 (February 1, 2009): 333–50. http://dx.doi.org/10.1175/2008jpo3954.1.

Full text
Abstract:
Abstract The structure of baroclinic vortices generated by horizontal flow separation past a sloping headland in deep, stably stratified waters is investigated. The most distinctive feature of these eddies is that their cores are strongly tilted with respect to the stratification, yet their velocity field remains quasi-horizontal. Field observations and numerical simulations are used to explore the consequences of the strong tilt on the eddy baroclinic structure. It is found that the background density field is altered in such a way as to maintain a pressure minimum in the tilted vortex cores. This adjustment results in a fundamental asymmetry of the density field. Isopycnals are deflected upward on the shoreward side and downward on the opposite side of the eddy center. The resulting pattern closely resembles the asymmetries of azimuthal wavenumber one that develop when tropical cyclones become tilted by an environmental shear. The authors provide a simple analytical model that suggests this structure is obtained via a balance between the centrifugal force and the horizontal pressure gradient. As the eddies release from the boundary, adjust, and decay, their tilt as well as the associated density perturbation decrease and lose coherence. It is suggested that this may lead to a conversion of potential energy into kinetic energy.
APA, Harvard, Vancouver, ISO, and other styles
10

TIPPETT, MICHAEL K. "Transient moist baroclinic instability." Tellus A 51, no. 2 (March 1999): 273–88. http://dx.doi.org/10.1034/j.1600-0870.1999.t01-2-00008.x.

Full text
APA, Harvard, Vancouver, ISO, and other styles
11

Tippett, Michael K. "Transient moist baroclinic instability." Tellus A: Dynamic Meteorology and Oceanography 51, no. 2 (January 1999): 273–88. http://dx.doi.org/10.3402/tellusa.v51i2.12321.

Full text
APA, Harvard, Vancouver, ISO, and other styles
12

Castrejón-Pita, A. A., and P. L. Read. "Synchronization in baroclinic systems." Journal of Physics: Conference Series 137 (November 1, 2008): 012016. http://dx.doi.org/10.1088/1742-6596/137/1/012016.

Full text
APA, Harvard, Vancouver, ISO, and other styles
13

McTaggart-Cowan, Ron, John R. Gyakum, and Richard W. Moore. "The Baroclinic Moisture Flux." Monthly Weather Review 145, no. 1 (December 14, 2016): 25–47. http://dx.doi.org/10.1175/mwr-d-16-0153.1.

Full text
Abstract:
Abstract As subsaturated air ascends sloping isentropic surfaces, adiabatic expansion results in cooling and relative moistening. This process is an effective way to precondition the atmosphere for efficient moist processes while bringing parcels to saturation, and thereafter acts to maintain saturation during condensation. The goal of this study is to develop a diagnostic quantity that highlights circulations and regions in which the process of parcel moistening by isentropic ascent is active. Among the many features that rely on this process for the generation of an important fraction of their energy are oceanic cyclones, transitioning tropical cyclones, warm conveyor belts, diabatic Rossby vortices, and predecessor rain events. The baroclinic moisture flux (BMF) is defined as moisture transport by the component of vertical motion associated with isentropic upgliding. In warm conveyor belt and diabatic Rossby vortex case studies, the BMF appears to be successful in identifying the portion of the circulation in which this process is actively bringing parcels to saturation to promote the formation of clouds and precipitation. On a broader scale, the climatological maxima of the BMF highlight regions in which parcel moistening by isentropic ascent is anticipated to have a nonnegligible impact on the atmospheric state either through the action of the mean flow or via the repeated occurrence of isolated large-BMF events. The process-centric foundation of the BMF makes it useful as a filtering or exploratory variable, with the potential for extension into predictive applications.
APA, Harvard, Vancouver, ISO, and other styles
14

Miller, Timothy L., and Basil N. Antar. "Viscous Nongeostrophic Baroclinic Instability." Journal of the Atmospheric Sciences 43, no. 4 (February 1986): 329–38. http://dx.doi.org/10.1175/1520-0469(1986)043<0329:vnbi>2.0.co;2.

Full text
APA, Harvard, Vancouver, ISO, and other styles
15

Snyder, Christopher M., and Richard S. Lindzen. "Upper-Level Baroclinic Instability." Journal of the Atmospheric Sciences 45, no. 17 (September 1988): 2445–59. http://dx.doi.org/10.1175/1520-0469(1988)045<2445:ulbi>2.0.co;2.

Full text
APA, Harvard, Vancouver, ISO, and other styles
16

Pedlosky, J., and R. M. Samelson. "Radiation-induced baroclinic instability." Geophysical & Astrophysical Fluid Dynamics 58, no. 1-4 (July 1991): 243–62. http://dx.doi.org/10.1080/03091929108227341.

Full text
APA, Harvard, Vancouver, ISO, and other styles
17

Appleby, J. C. "Selection of baroclinic waves." Quarterly Journal of the Royal Meteorological Society 114, no. 482 (July 1988): 1173–79. http://dx.doi.org/10.1002/qj.49711448214.

Full text
APA, Harvard, Vancouver, ISO, and other styles
18

Zhang, Xuehong, Qingcun Zeng, and Ning Bao. "Nonlinear baroclinic haurwitz waves." Advances in Atmospheric Sciences 3, no. 3 (August 1986): 330–40. http://dx.doi.org/10.1007/bf02678653.

Full text
APA, Harvard, Vancouver, ISO, and other styles
19

Thompson, A., L. Stefanova, and T. N. Krishnamurti. "Baroclinic splitting of jets." Meteorology and Atmospheric Physics 100, no. 1-4 (August 2008): 257–74. http://dx.doi.org/10.1007/s00703-008-0308-5.

Full text
APA, Harvard, Vancouver, ISO, and other styles
20

Meachem, Stephen. "Non-modal baroclinic instability." Dynamics of Atmospheres and Oceans 12, no. 1 (October 1988): 19–45. http://dx.doi.org/10.1016/0377-0265(88)90013-9.

Full text
APA, Harvard, Vancouver, ISO, and other styles
21

von Larcher, Th, W. Beyer, and C. Egbers. "Experiments on baroclinic instabilities." PAMM 4, no. 1 (December 2004): 504–5. http://dx.doi.org/10.1002/pamm.200410233.

Full text
APA, Harvard, Vancouver, ISO, and other styles
22

Brink, K. H., and J. Pedlosky. "The Structure of Baroclinic Modes in the Presence of Baroclinic Mean Flow." Journal of Physical Oceanography 50, no. 1 (January 2020): 239–53. http://dx.doi.org/10.1175/jpo-d-19-0123.1.

Full text
Abstract:
AbstractThis contribution seeks to understand the vertical structure of linearized quasigeostrophic baroclinic modes when they are modified by the presence of a baroclinic mean flow and associated potential vorticity gradients. It is found that even modest, O(0.05 m s−1), mean flows can give rise to very substantial changes in modal structures, often in the sense of increased surface intensification. The extent to which stable modes are modified depends strongly on the direction of Rossby wave propagation. Further, baroclinically unstable solutions can appear, and a meaningful inviscid critical-layer solution can occur at the transition to instability when the horizontal gradient of potential vorticity changes sign at some depth within the water column. In addition, the gravest, n = 0, vertical stable mode is no longer strictly barotropic, but rather it can carry density variability at frequencies much higher than those possible for baroclinic (higher) Rossby wave modes. This finding appears to be consistent with oceanic current-meter observations that suggest temperature variability propagation even when the frequency is too high for traditional baroclinic Rossby waves to exist.
APA, Harvard, Vancouver, ISO, and other styles
23

Williams, Paul D., and Christopher W. Kelsall. "The Dynamics of Baroclinic Zonal Jets*." Journal of the Atmospheric Sciences 72, no. 3 (February 24, 2015): 1137–51. http://dx.doi.org/10.1175/jas-d-14-0027.1.

Full text
Abstract:
Abstract Multiple alternating zonal jets are a ubiquitous feature of planetary atmospheres and oceans. However, most studies to date have focused on the special case of barotropic jets. Here, the dynamics of freely evolving baroclinic jets are investigated using a two-layer quasigeostrophic annulus model with sloping topography. In a suite of 15 numerical simulations, the baroclinic Rossby radius and baroclinic Rhines scale are sampled by varying the stratification and root-mean-square eddy velocity, respectively. Small-scale eddies in the initial state evolve through geostrophic turbulence and accelerate zonally as they grow in horizontal scale, first isotropically and then anisotropically. This process leads ultimately to the formation of jets, which take about 2500 rotation periods to equilibrate. The kinetic energy spectrum of the equilibrated baroclinic zonal flow steepens from a −3 power law at small scales to a −5 power law near the jet scale. The conditions most favorable for producing multiple alternating baroclinic jets are large baroclinic Rossby radius (i.e., strong stratification) and small baroclinic Rhines scale (i.e., weak root-mean-square eddy velocity). The baroclinic jet width is diagnosed objectively and found to be 2.2–2.8 times larger than the baroclinic Rhines scale, with a best estimate of 2.5 times larger. This finding suggests that Rossby wave motions must be moving at speeds of approximately 6 times the turbulent eddy velocity in order to be capable of arresting the isotropic inverse energy cascade.
APA, Harvard, Vancouver, ISO, and other styles
24

Boljka, Lina, Theodore G. Shepherd, and Michael Blackburn. "On the Coupling between Barotropic and Baroclinic Modes of Extratropical Atmospheric Variability." Journal of the Atmospheric Sciences 75, no. 6 (May 18, 2018): 1853–71. http://dx.doi.org/10.1175/jas-d-17-0370.1.

Full text
Abstract:
Abstract The baroclinic and barotropic components of atmospheric dynamics are usually viewed as interlinked through the baroclinic life cycle, with baroclinic growth of eddies connected to heat fluxes, barotropic decay connected to momentum fluxes, and the two eddy fluxes connected through the Eliassen–Palm wave activity. However, recent observational studies have suggested that these two components of the dynamics are largely decoupled in their variability, with variations in the zonal mean flow associated mainly with the momentum fluxes, variations in the baroclinic wave activity associated mainly with the heat fluxes, and essentially no correlation between the two. These relationships are examined in a dry dynamical core model under different configurations and in Southern Hemisphere observations, considering different frequency bands to account for the different time scales of atmospheric variability. It is shown that at intermediate periods longer than 10 days, the decoupling of the baroclinic and barotropic modes of variability can indeed occur as the eddy kinetic energy at those time scales is only affected by the heat fluxes and not the momentum fluxes. The baroclinic variability includes the oscillator model with periods of 20–30 days. At both the synoptic time scale and the quasi-steady limit, the baroclinic and barotropic modes of variability are linked, consistent with baroclinic life cycles and the positive baroclinic feedback mechanism, respectively. In the quasi-steady limit, the pulsating modes of variability and their correlations depend sensitively on the model climatology.
APA, Harvard, Vancouver, ISO, and other styles
25

Li, Xinyu, Riyu Lu, Richard J. Greatbatch, Gen Li, and Xiaowei Hong. "Maintenance Mechanism for the Teleconnection Pattern over the High Latitudes of the Eurasian Continent in Summer." Journal of Climate 33, no. 3 (February 1, 2020): 1017–30. http://dx.doi.org/10.1175/jcli-d-19-0362.1.

Full text
Abstract:
AbstractThere is a zonally oriented teleconnection pattern over the high-latitude Eurasian continent, which is maintained through baroclinic energy conversion. In this study, we investigate the unique features of the maintenance mechanism of this teleconnection. It is found that the baroclinic energy conversion is most efficient in both the midtroposphere and the lower troposphere, and that the baroclinic energy conversion in the lower troposphere is comparable to that in the midtroposphere. Further results indicate that the basic state plays a crucial role in the baroclinic energy conversion. For both the middle and lower troposphere, the atmospheric stability is low and the Coriolis parameter is large over high-latitude Eurasia, favoring strong baroclinic energy conversion. Particularly, in the lower troposphere, the atmospheric stability exhibits a clear land–sea contrast, favoring baroclinic energy conversion over the continents rather than the oceans. Furthermore, in the lower troposphere, the in-phase configuration of the meridional wind and temperature anomalies, which results from the strong meridional gradient of mean temperature around the north edge of the Eurasian continent, also significantly contributes to baroclinic energy conversion. This study highlights the role of the basic state of temperature rather than zonal wind in maintaining the high-latitude teleconnection through baroclinic energy conversion.
APA, Harvard, Vancouver, ISO, and other styles
26

Musgrave, R. C. "Energy Fluxes in Coastal Trapped Waves." Journal of Physical Oceanography 49, no. 12 (December 2019): 3061–68. http://dx.doi.org/10.1175/jpo-d-18-0172.1.

Full text
Abstract:
AbstractThe calculation of energy flux in coastal trapped wave modes is reviewed in the context of tidal energy pathways near the coast. The significant barotropic pressures and currents associated with coastal trapped wave modes mean that large errors in estimating the wave flux are incurred if only the baroclinic component is considered. A specific example is given showing that baroclinic flux constitutes only 10% of the flux in a mode-1 wave for a reasonable choice of stratification and bathymetry. The interpretation of baroclinic energy flux and barotropic-to-baroclinic conversion at the coast is discussed: in contrast to the open ocean, estimates of baroclinic energy flux do not represent a wave energy flux; neither does conversion represent the scattering of energy from the tidal Kelvin wave to higher modes.
APA, Harvard, Vancouver, ISO, and other styles
27

Swaters, Gordon E. "Modal Interpretation for the Ekman Destabilization of Inviscidly Stable Baroclinic Flow in the Phillips Model." Journal of Physical Oceanography 40, no. 4 (April 1, 2010): 830–39. http://dx.doi.org/10.1175/2009jpo4311.1.

Full text
Abstract:
Abstract Ekman boundary layers can lead to the destabilization of baroclinic flow in the Phillips model that, in the absence of dissipation, is nonlinearly stable in the sense of Liapunov. It is shown that the Ekman-induced instability of inviscidly stable baroclinic flow in the Phillips model occurs if and only if the kinematic phase velocity associated with the dissipation lies outside the interval bounded by the greatest and least neutrally stable Rossby wave phase velocities. Thus, Ekman-induced destabilization does not correspond to a coalescence of the barotropic and baroclinic Rossby modes as in classical inviscid baroclinic instability. The differing modal mechanisms between the two instability processes is the reason why subcritical baroclinic shears in the classical theory can be destabilized by an Ekman layer, even in the zero dissipation limit of the theory.
APA, Harvard, Vancouver, ISO, and other styles
28

Wu, Li Dan, and Chun Bao Miao. "The Mechanism of Internal Tide Generation over Weak Topography." Advanced Materials Research 588-589 (November 2012): 1972–78. http://dx.doi.org/10.4028/www.scientific.net/amr.588-589.1972.

Full text
Abstract:
Here the process of internal tide generation over idealized (sinusoidal) topography was investigated using numerical techniques, in which the barotropic-to-baroclinic energy conversion was discussed. The result shows that when the wavenumber of the sinusoidal topography, ktopo, is equal to the horizontal wavenumber of the m-th baroclinic mode km, the conversion from the barotropic tide to the m-th baroclinic mode is enhanced with the increase of topography length; When the wavenumber of the sinusoidal topography is not equal to horizontal wavenumbers of any baroclinic modes, ktopo≠km(m=1,2,...), conversion from the barotropic energy to baroclinic modes is decreased with the increase of topography length. Furthermore, it shows that in resonance case, the phase of the perturbation pressure gradually agrees with the phase of the truncated sinusoidal topography, and the conversion rate is always positive over the topography, thus the baroclinic mode which matches with the wavenumber of the sinusoidal topography persists in absorbing energy from the barotropic tide, and the conversion rate is increased.
APA, Harvard, Vancouver, ISO, and other styles
29

Skyllingstad, Eric D., and R. M. Samelson. "Baroclinic Frontal Instabilities and Turbulent Mixing in the Surface Boundary Layer. Part I: Unforced Simulations." Journal of Physical Oceanography 42, no. 10 (June 1, 2012): 1701–16. http://dx.doi.org/10.1175/jpo-d-10-05016.1.

Full text
Abstract:
Abstract Interaction between mixed layer baroclinic eddies and small-scale turbulence is studied using a nonhydrostatic large-eddy simulation (LES) model. Free, unforced flow evolution is considered, for a standard initialization consisting of an 80-m-deep mixed layer with a superposed warm filament and two frontal interfaces in geostrophic balance, on a model domain roughly 5 km × 10 km × 120 m, with an isotropic 3-m computational grid. Results from these unforced experiments suggest that shear generated in narrow frontal zones can support weak three-dimensional turbulence that is directly linked to the larger-scale baroclinic waves. Two separate but closely related issues are addressed: 1) the possible development of enhanced turbulent mixing associated with the baroclinic wave activity and 2) the existence of a downscale transfer of energy from the baroclinic wave scale to the turbulent dissipation scale. The simulations show enhanced turbulence associated with the baroclinic waves and enhanced turbulent heat flux across the isotherms of the imposed frontal boundary, relative to background levels. This turbulence develops on isolated small-scale frontal features that form as the result of frontogenetic processes operating on the baroclinic wave scale and not as the result of a continuous, inertial forward cascade through the intermediate scales. Analysis of the spectrally decomposed kinetic energy budget indicates that large-scale baroclinic eddy energy is directly transferred to small-scale turbulence, with weaker forcing at intermediate scales. For fronts with significant baroclinic wave activity, cross-frontal eddy fluxes computed from correlations of fluctuations from means along the large-scale frontal axis generally agreed with simple theoretical estimates.
APA, Harvard, Vancouver, ISO, and other styles
30

LaCasce, J. H. "Surface Quasigeostrophic Solutions and Baroclinic Modes with Exponential Stratification." Journal of Physical Oceanography 42, no. 4 (April 1, 2012): 569–80. http://dx.doi.org/10.1175/jpo-d-11-0111.1.

Full text
Abstract:
Abstract The author derives baroclinic modes and surface quasigeostrophic (SQG) solutions with exponential stratification and compares the results to those obtained with constant stratification. The SQG solutions with exponential stratification decay more rapidly in the vertical and have weaker near-surface velocities. This then compounds the previously noted problem that SQG underpredicts the velocities associated with a given surface density anomaly. The author also examines how the SQG solutions project onto the baroclinic modes. With constant stratification, SQG waves larger than deformation scale project primarily onto the barotropic mode and to a lesser degree onto the first baroclinic mode. However, with exponential stratification, the largest projection is on the first baroclinic mode. The effect is even more pronounced over rough bottom topography. Therefore, large-scale SQG waves will look like the first baroclinic mode and vice versa, with realistic stratification.
APA, Harvard, Vancouver, ISO, and other styles
31

Kurgansky, Michael V. "Helicity production and maintenance in a baroclinic atmosphere." Meteorologische Zeitschrift 15, no. 4 (August 23, 2006): 409–16. http://dx.doi.org/10.1127/0941-2948/2006/0148.

Full text
APA, Harvard, Vancouver, ISO, and other styles
32

Egger, Joseph. "Baroclinic instability in the two-layer model: interpretations." Meteorologische Zeitschrift 18, no. 5 (October 1, 2009): 559–65. http://dx.doi.org/10.1127/0941-2948/2009/0405.

Full text
APA, Harvard, Vancouver, ISO, and other styles
33

Garrett, Chris, and Theo Gerkema. "On the Body-Force Term in Internal-Tide Generation." Journal of Physical Oceanography 37, no. 8 (August 1, 2007): 2172–75. http://dx.doi.org/10.1175/jpo3165.1.

Full text
Abstract:
Abstract The generation of internal tides can be ascribed to the action of a buoyancy force caused by the flow of the barotropic tide over topographic features. It is commonly assumed that the barotropic flow can be taken as hydrostatic, but it is shown here that this leads to a linearized governing equation for the baroclinic tide that is only valid if the baroclinic tide is also hydrostatic. A governing equation for the baroclinic tide, valid for any situation, is derived here and is shown to be exactly equivalent to a simple transformation of the governing equation for the combined barotropic and baroclinic tides.
APA, Harvard, Vancouver, ISO, and other styles
34

Teubler, Franziska, and Michael Riemer. "Potential-vorticity dynamics of troughs and ridges within Rossby wave packets during a 40-year reanalysis period." Weather and Climate Dynamics 2, no. 3 (July 6, 2021): 535–59. http://dx.doi.org/10.5194/wcd-2-535-2021.

Full text
Abstract:
Abstract. Rossby wave packets (RWPs) are fundamental to midlatitude dynamics and govern weather systems from their individual life cycles to their climatological distributions. Renewed interest in RWPs as precursors to high-impact weather events and in the context of atmospheric predictability motivates this study to revisit the dynamics of RWPs. A quantitative potential-vorticity (PV) framework is employed. Based on the well-established PV thinking of midlatitude dynamics, the processes governing RWP amplitude evolution comprise group propagation of Rossby waves, baroclinic interaction, the impact of upper-tropospheric divergent flow, and direct diabatic PV modification by nonconservative processes. An advantage of the PV framework is that the impact of moist processes is more directly diagnosed than in alternative, established frameworks for RWP dynamics. The mean dynamics of more than 6000 RWPs from 1979–2017 are presented using ERA5 data, complemented with nonconservative tendencies from the Year of Tropical Convection data (available 2008–2010). Confirming a pre-existing model of RWP dynamics, group propagation within RWPs is consistent with linear barotropic theory, and baroclinic and divergent amplifications occur most prominently during the mature stage and towards the trailing edge of RWPs. Refining the pre-existing model, the maximum of divergent amplification occurs in advance of maximum baroclinic growth, and baroclinic interaction tends to weaken RWP amplitude towards the leading edge. “Downstream baroclinic development” is confirmed to provide a valid description of RWP dynamics in both summer and winter, although baroclinic growth is substantially smaller (about 50 %) in summer. Longwave radiative cooling makes a first-order contribution to ridge and trough amplitude, with the potential that this contribution is partly associated with cloud-radiative effects. The direct impact of other nonconservative tendencies, including latent heat release, is an order of magnitude smaller than longwave radiative cooling. Arguably, latent heat release still has a substantial impact on RWPs by invigorating upper-tropospheric divergence. The divergent flow amplifies ridges and weakens troughs. This impact is of leading order and larger than that of baroclinic growth. To the extent that divergence is associated with latent heat release below, our results show that moist processes contribute to the well-known asymmetry in the spatial scale of troughs and ridges. For ridges, divergent amplification is strongly coupled to baroclinic growth and enhanced latent heat release. We thus propose that the life cycle of ridges is best described in terms of downstream moist-baroclinic development. Consistent with theories of moist-baroclinic instability, both the amplitude and the relative location of latent heat release within the developing wave pattern depend on the state of the baroclinic development. Taking this “phasing” aspect into account, we provide some evidence that variability in the strength of divergent ridge amplification can predominantly be attributed to variability in latent heat release below rather than to secondary circulations associated with the dry dynamics of a baroclinic wave.
APA, Harvard, Vancouver, ISO, and other styles
35

Hoffman, Eric G. "Surface Potential Temperature as an Analysis and Forecasting Tool." Meteorological Monographs 55 (November 1, 2008): 163–82. http://dx.doi.org/10.1175/0065-9401-33.55.163.

Full text
Abstract:
Abstract In the last decade, Fred Sanders was often critical of current surface analysis techniques. This led to his promoting the use of surface potential temperatures to distinguish between fronts, baroclinic troughs, and non-frontal baroclinic zones, and to the development of a climatology of surface baroclinic zones. In this paper, criticisms of current surface analysis techniques and the usefulness of surface potential temperature analyses are discussed. Case examples are used to compare potential temperature analyses and current National Centers for Environmental Prediction analyses. The 1-yr climatology of Sanders and Hoffman is reconstructed using a composite technique. Annual and seasonal mean potential temperature analyses over the continental United States, southern Canada, northern Mexico, and adjacent coastal waters are presented. In addition, gridpoint frequencies of moderate and strong potential temperature gradients are calculated. The results of the mean potential temperature analyses show that moderate and strong surface baroclinic zones are favored along the coastlines and the slopes of the North American cordillera. Additional subsynoptic details, not found in Sanders and Hoffman, are identified. The availability of the composite results allows for the calculation of potential temperature gradient anomalies. It is shown that these anomalies can be used to identify significant frontal baroclinic zones that are associated with weak potential temperature gradients. Together the results and reviews in this paper show that surface potential temperature analyses are a valuable forecasting and analysis tool allowing analysts to distinguish and identify fronts, baroclinic troughs, and nonfrontal baroclinic zones.
APA, Harvard, Vancouver, ISO, and other styles
36

Kazbekov, Askar, Keishi Kumashiro, and Adam M. Steinberg. "Enstrophy transport in swirl combustion." Journal of Fluid Mechanics 876 (August 6, 2019): 715–32. http://dx.doi.org/10.1017/jfm.2019.551.

Full text
Abstract:
The contributions of vortex stretching, dilatation, baroclinic torque and viscous diffusion to Reynolds-averaged enstrophy transport in turbulent swirl flames were experimentally measured using tomographic particle image velocimetry and $\text{CH}_{2}\text{O}$ planar laser induced fluorescence at jet Reynolds numbers of 26 000–51 000. The mean baroclinic torque was determined by subtracting the other terms in the enstrophy transport equation from the mean Lagrangian derivative. Enstrophy production from baroclinic torque was found to be significant relative to the other transport terms across all conditions studies. This result contrasts with direct numerical simulations of flames in homogeneous isotropic turbulence, which show a decreasing relative significance of baroclinic torque with increasing turbulence intensity (e.g. Bobbitt, Lapointe & Blanquart, Phys. Fluids, vol. 28 (1), 2016, 015101). Hence, the significance of baroclinic enstrophy production in flames is not determined entirely by the local turbulence and flame properties, but also depends on the configuration-specific pressure field.
APA, Harvard, Vancouver, ISO, and other styles
37

Sitte, B., and C. Egbers. "LDV-measurements on baroclinic waves." Physics and Chemistry of the Earth, Part B: Hydrology, Oceans and Atmosphere 24, no. 5 (January 1999): 473–76. http://dx.doi.org/10.1016/s1464-1909(99)00031-3.

Full text
APA, Harvard, Vancouver, ISO, and other styles
38

Spall, Michael A. "Baroclinic Jets in Confluent Flow*." Journal of Physical Oceanography 27, no. 6 (June 1997): 1054–71. http://dx.doi.org/10.1175/1520-0485(1997)027<1054:bjicf>2.0.co;2.

Full text
APA, Harvard, Vancouver, ISO, and other styles
39

Walker, Alison, and Joseph Pedlosky. "Instability of Meridional Baroclinic Currents*." Journal of Physical Oceanography 32, no. 3 (March 2002): 1075–93. http://dx.doi.org/10.1175/1520-0485(2002)032<1075:iombc>2.0.co;2.

Full text
APA, Harvard, Vancouver, ISO, and other styles
40

Lesieur †, Marcel, Olivier Métais, and Elodie Garnier. "Baroclinic instability and severe storms." Journal of Turbulence 1 (January 2000): N2. http://dx.doi.org/10.1088/1468-5248/1/1/002.

Full text
APA, Harvard, Vancouver, ISO, and other styles
41

Reinaud, Jean N. "Baroclinic toroidal quasi-geostrophic vortices." Physics of Fluids 32, no. 5 (May 1, 2020): 056601. http://dx.doi.org/10.1063/5.0005942.

Full text
APA, Harvard, Vancouver, ISO, and other styles
42

Zuo, Da-Wei, Yi-Tian Gao, Yu-Jie Feng, Long Xue, and Yu-Hao Sun. "Rogue waves in baroclinic flows." Theoretical and Mathematical Physics 191, no. 2 (May 2017): 725–37. http://dx.doi.org/10.1134/s0040577917050129.

Full text
APA, Harvard, Vancouver, ISO, and other styles
43

Farrell, Brian F. "Optimal Excitation of Baroclinic Waves." Journal of the Atmospheric Sciences 46, no. 9 (May 1989): 1193–206. http://dx.doi.org/10.1175/1520-0469(1989)046<1193:oeobw>2.0.co;2.

Full text
APA, Harvard, Vancouver, ISO, and other styles
44

Staley, D. O. "Ageostrophic Subsynoptic-Scale Baroclinic Instability." Journal of the Atmospheric Sciences 46, no. 19 (October 1989): 3065–68. http://dx.doi.org/10.1175/1520-0469(1989)046<3065:assbi>2.0.co;2.

Full text
APA, Harvard, Vancouver, ISO, and other styles
45

Reinhold, Brian. "Orographic Modulation of Baroclinic Instability." Journal of the Atmospheric Sciences 47, no. 14 (July 1990): 1697–713. http://dx.doi.org/10.1175/1520-0469(1990)047<1697:omobi>2.0.co;2.

Full text
APA, Harvard, Vancouver, ISO, and other styles
46

Staley, D. O. "Baroclinic Instability and Isentropic Slope." Journal of the Atmospheric Sciences 48, no. 9 (May 1991): 1133–40. http://dx.doi.org/10.1175/1520-0469(1991)048<1133:biais>2.0.co;2.

Full text
APA, Harvard, Vancouver, ISO, and other styles
47

Pedlosky, Joseph. "Baroclinic Instability Localized by Dissipation." Journal of the Atmospheric Sciences 49, no. 13 (July 1992): 1161–70. http://dx.doi.org/10.1175/1520-0469(1992)049<1161:bilbd>2.0.co;2.

Full text
APA, Harvard, Vancouver, ISO, and other styles
48

Lindzen, Richards S. "Baroclinic Neutrality and the Tropopause." Journal of the Atmospheric Sciences 50, no. 8 (April 1993): 1148–51. http://dx.doi.org/10.1175/1520-0469(1993)050<1148:bnatt>2.0.co;2.

Full text
APA, Harvard, Vancouver, ISO, and other styles
49

Farrell, Brian F., and Petros J. Ioannou. "Stochastic Dynamics of Baroclinic Waves." Journal of the Atmospheric Sciences 50, no. 24 (December 1993): 4044–57. http://dx.doi.org/10.1175/1520-0469(1993)050<4044:sdobw>2.0.co;2.

Full text
APA, Harvard, Vancouver, ISO, and other styles
50

Neven, E. C. "Baroclinic Modons on a Sphere." Journal of the Atmospheric Sciences 51, no. 11 (June 1994): 1447–64. http://dx.doi.org/10.1175/1520-0469(1994)051<1447:bmoas>2.0.co;2.

Full text
APA, Harvard, Vancouver, ISO, and other styles
We offer discounts on all premium plans for authors whose works are included in thematic literature selections. Contact us to get a unique promo code!

To the bibliography