Maximum potential intensity

Last updated

The maximum potential intensity of a tropical cyclone is the theoretical limit of the strength of a tropical cyclone.

Contents

Maximum potential intensity

Due to surface friction, the inflow only partially conserves angular momentum. Thus, the sea surface lower boundary acts as both a source (evaporation) and sink (friction) of energy for the system. This fact leads to the existence of a theoretical upper bound on the strongest wind speed that a tropical cyclone can attain. Because evaporation increases linearly with wind speed (just as climbing out of a pool feels much colder on a windy day), there is a positive feedback on energy input into the system known as the Wind-Induced Surface Heat Exchange (WISHE) feedback. [1] This feedback is offset when frictional dissipation, which increases with the cube of the wind speed, becomes sufficiently large. This upper bound is called the "maximum potential intensity", , and is given by

where is the temperature of the sea surface, is the temperature of the outflow ([K]), is the enthalpy difference between the surface and the overlying air ([J/kg]), and and are the surface exchange coefficients (dimensionless) of enthalpy and momentum, respectively. [2] The surface-air enthalpy difference is taken as , where is the saturation enthalpy of air at sea surface temperature and sea-level pressure and is the enthalpy of boundary layer air overlying the surface.

The maximum potential intensity is predominantly a function of the background environment alone (i.e. without a tropical cyclone), and thus this quantity can be used to determine which regions on Earth can support tropical cyclones of a given intensity, and how these regions may evolve in time. [3] [4] Specifically, the maximum potential intensity has three components, but its variability in space and time is due predominantly to the variability in the surface-air enthalpy difference component .

Derivation

A tropical cyclone may be viewed as a heat engine that converts input heat energy from the surface into mechanical energy that can be used to do mechanical work against surface friction. At equilibrium, the rate of net energy production in the system must equal the rate of energy loss due to frictional dissipation at the surface, i.e.

The rate of energy loss per unit surface area from surface friction, , is given by

where is the density of near-surface air ([kg/m3]) and is the near surface wind speed ([m/s]).

The rate of energy production per unit surface area, is given by

where is the heat engine efficiency and is the total rate of heat input into the system per unit surface area. Given that a tropical cyclone may be idealized as a Carnot heat engine, the Carnot heat engine efficiency is given by

Heat (enthalpy) per unit mass is given by

where is the heat capacity of air, is air temperature, is the latent heat of vaporization, and is the concentration of water vapor. The first component corresponds to sensible heat and the second to latent heat.

There are two sources of heat input. The dominant source is the input of heat at the surface, primarily due to evaporation. The bulk aerodynamic formula for the rate of heat input per unit area at the surface, , is given by

where represents the enthalpy difference between the ocean surface and the overlying air. The second source is the internal sensible heat generated from frictional dissipation (equal to ), which occurs near the surface within the tropical cyclone and is recycled to the system.

Thus, the total rate of net energy production per unit surface area is given by

Setting and taking (i.e. the rotational wind speed is dominant) leads to the solution for given above. This derivation assumes that total energy input and loss within the system can be approximated by their values at the radius of maximum wind. The inclusion of acts to multiply the total heat input rate by the factor . Mathematically, this has the effect of replacing with in the denominator of the Carnot efficiency.

An alternative definition for the maximum potential intensity, which is mathematically equivalent to the above formulation, is

where CAPE stands for the Convective Available Potential Energy, is the CAPE of an air parcel lifted from saturation at sea level in reference to the environmental sounding, is the CAPE of the boundary layer air, and both quantities are calculated at the radius of maximum wind. [5]

Characteristic values and variability on Earth

On Earth, a characteristic temperature for is 300 K and for is 200 K, corresponding to a Carnot efficiency of . The ratio of the surface exchange coefficients, , is typically taken to be 1. However, observations suggest that the drag coefficient varies with wind speed and may decrease at high wind speeds within the boundary layer of a mature hurricane. [6] Additionally, may vary at high wind speeds due to the effect of sea spray on evaporation within the boundary layer. [7]

A characteristic value of the maximum potential intensity, , is 80 metres per second (180 mph; 290 km/h). However, this quantity varies significantly across space and time, particularly within the seasonal cycle, spanning a range of 0 to 100 metres per second (0 to 224 mph; 0 to 360 km/h). [5] This variability is primarily due to variability in the surface enthalpy disequilibrium ( ) as well as in the thermodynamic structure of the troposphere, which are controlled by the large-scale dynamics of the tropical climate. These processes are modulated by factors including the sea surface temperature (and underlying ocean dynamics), background near-surface wind speed, and the vertical structure of atmospheric radiative heating. [8] The nature of this modulation is complex, particularly on climate time-scales (decades or longer). On shorter time-scales, variability in the maximum potential intensity is commonly linked to sea surface temperature perturbations from the tropical mean, as regions with relatively warm water have thermodynamic states much more capable of sustaining a tropical cyclone than regions with relatively cold water. [9] However, this relationship is indirect via the large-scale dynamics of the tropics; the direct influence of the absolute sea surface temperature on is weak in comparison.

Empirical limit

An empirical limit on tropical cyclone intensity can also be computed using the following formula:

Where is the maximum potential velocity in meters per second; is the sea surface temperature underneath the center of the tropical cyclone, is a reference temperature (30 ˚C) and , and are curve-fit constants. When , , and , the graph generated by this function corresponds to the 99th percentile of empirical tropical cyclone intensity data. [10]

See also

Related Research Articles

<span class="mw-page-title-main">Enthalpy</span> Measure of energy in a thermodynamic system

In thermodynamics, enthalpy, is the sum of a thermodynamic system's internal energy and the product of its pressure and volume. It is a state function used in many measurements in chemical, biological, and physical systems at a constant pressure, which is conveniently provided by the large ambient atmosphere. The pressure–volume term expresses the work required to establish the system's physical dimensions, i.e. to make room for it by displacing its surroundings. The pressure-volume term is very small for solids and liquids at common conditions, and fairly small for gases. Therefore, enthalpy is a stand-in for energy in chemical systems; bond, lattice, solvation, and other chemical "energies" are actually enthalpy differences. As a state function, enthalpy depends only on the final configuration of internal energy, pressure, and volume, not on the path taken to achieve it.

In fluid mechanics, the Grashof number is a dimensionless number which approximates the ratio of the buoyancy to viscous forces acting on a fluid. It frequently arises in the study of situations involving natural convection and is analogous to the Reynolds number.

Conduction is the process by which heat is transferred from the hotter end to the colder end of an object. The ability of the object to conduct heat is known as its thermal conductivity, and is denoted k.

<span class="mw-page-title-main">First law of thermodynamics</span> Law of thermodynamics establishing the conservation of energy

The first law of thermodynamics is a formulation of the law of conservation of energy in the context of thermodynamic processes in which two principle forms of energy transfer, heat and thermodynamic work, are distinguished that modify a thermodynamic system of a constant amount of matter. The law also defines the internal energy of a system, an extensive property for taking account of the balance of these energies in the system. Energy cannot be created or destroyed, but it can be transformed from one form to another. In an isolated system the sum of all forms of energy is constant.

<span class="mw-page-title-main">Heat equation</span> Partial differential equation describing the evolution of temperature in a region

In mathematics and physics, the heat equation is a certain partial differential equation. Solutions of the heat equation are sometimes known as caloric functions. The theory of the heat equation was first developed by Joseph Fourier in 1822 for the purpose of modeling how a quantity such as heat diffuses through a given region.

<span class="mw-page-title-main">Heat capacity</span> Physical property describing the energy required to change a materials temperature

Heat capacity or thermal capacity is a physical property of matter, defined as the amount of heat to be supplied to an object to produce a unit change in its temperature. The SI unit of heat capacity is joule per kelvin (J/K).

<span class="mw-page-title-main">Gibbs free energy</span> Type of thermodynamic potential; useful for calculating reversible work in certain systems

In thermodynamics, the Gibbs free energy is a thermodynamic potential that can be used to calculate the maximum amount of work, other than pressure-volume work, that may be performed by a thermodynamically closed system at constant temperature and pressure. It also provides a necessary condition for processes such as chemical reactions that may occur under these conditions. The Gibbs free energy is expressed as

The standard enthalpy of reaction for a chemical reaction is the difference between total product and total reactant molar enthalpies, calculated for substances in their standard states. The value can be approximately interpreted in terms of the total of the chemical bond energies for bonds broken and bonds formed.

<span class="mw-page-title-main">Isentropic process</span> Thermodynamic process that is reversible and adiabatic

In thermodynamics, an isentropic process is an idealized thermodynamic process that is both adiabatic and reversible. The work transfers of the system are frictionless, and there is no net transfer of heat or matter. Such an idealized process is useful in engineering as a model of and basis of comparison for real processes. This process is idealized because reversible processes do not occur in reality; thinking of a process as both adiabatic and reversible would show that the initial and final entropies are the same, thus, the reason it is called isentropic. Thermodynamic processes are named based on the effect they would have on the system. Even though in reality it is not necessarily possible to carry out an isentropic process, some may be approximated as such.

A continuity equation or transport equation is an equation that describes the transport of some quantity. It is particularly simple and powerful when applied to a conserved quantity, but it can be generalized to apply to any extensive quantity. Since mass, energy, momentum, electric charge and other natural quantities are conserved under their respective appropriate conditions, a variety of physical phenomena may be described using continuity equations.

In physics, the Einstein relation is a previously unexpected connection revealed independently by William Sutherland in 1904, Albert Einstein in 1905, and by Marian Smoluchowski in 1906 in their works on Brownian motion. The more general form of the equation in the classical case is

<span class="mw-page-title-main">Work (thermodynamics)</span> Type of energy transfer

Thermodynamic work is one of the principal processes by which a thermodynamic system can interact with its surroundings and exchange energy. This exchange results in externally measurable macroscopic forces on the system's surroundings, which can cause mechanical work, to lift a weight, for example, or cause changes in electromagnetic, or gravitational variables. The surroundings also can perform work on a thermodynamic system, which is measured by an opposite sign convention.

Ewald summation, named after Paul Peter Ewald, is a method for computing long-range interactions in periodic systems. It was first developed as the method for calculating the electrostatic energies of ionic crystals, and is now commonly used for calculating long-range interactions in computational chemistry. Ewald summation is a special case of the Poisson summation formula, replacing the summation of interaction energies in real space with an equivalent summation in Fourier space. In this method, the long-range interaction is divided into two parts: a short-range contribution, and a long-range contribution which does not have a singularity. The short-range contribution is calculated in real space, whereas the long-range contribution is calculated using a Fourier transform. The advantage of this method is the rapid convergence of the energy compared with that of a direct summation. This means that the method has high accuracy and reasonable speed when computing long-range interactions, and it is thus the de facto standard method for calculating long-range interactions in periodic systems. The method requires charge neutrality of the molecular system to accurately calculate the total Coulombic interaction. A study of the truncation errors introduced in the energy and force calculations of disordered point-charge systems is provided by Kolafa and Perram.

In fluid mechanics, potential vorticity (PV) is a quantity which is proportional to the dot product of vorticity and stratification. This quantity, following a parcel of air or water, can only be changed by diabatic or frictional processes. It is a useful concept for understanding the generation of vorticity in cyclogenesis, especially along the polar front, and in analyzing flow in the ocean.

<span class="mw-page-title-main">Heat</span> Type of energy transfer

In thermodynamics, heat is the thermal energy transferred between systems due to a temperature difference. In colloquial use, heat sometimes refers to thermal energy itself. Thermal energy is the kinetic energy of vibrating and colliding atoms in a substance.

<span class="mw-page-title-main">Diffusion</span> Transport of dissolved species from the highest to the lowest concentration region

Diffusion is the net movement of anything generally from a region of higher concentration to a region of lower concentration. Diffusion is driven by a gradient in Gibbs free energy or chemical potential. It is possible to diffuse "uphill" from a region of lower concentration to a region of higher concentration, as in spinodal decomposition. Diffusion is a stochastic process due to the inherent randomness of the diffusing entity and can be used to model many real-life stochastic scenarios. Therefore, diffusion and the corresponding mathematical models are used in several fields beyond physics, such as statistics, probability theory, information theory, neural networks, finance, and marketing.

<span class="mw-page-title-main">Entropy production</span> Development of entropy in a thermodynamic system

Entropy production is the amount of entropy which is produced during heat process to evaluate the efficiency of the process.

CFD stands for computational fluid dynamics. As per this technique, the governing differential equations of a flow system or thermal system are known in the form of Navier–Stokes equations, thermal energy equation and species equation with an appropriate equation of state. In the past few years, CFD has been playing an increasingly important role in building design, following its continuing development for over a quarter of a century. The information provided by CFD can be used to analyse the impact of building exhausts to the environment, to predict smoke and fire risks in buildings, to quantify indoor environment quality, and to design natural ventilation systems.

An axial fan is a type of fan that causes gas to flow through it in an axial direction, parallel to the shaft about which the blades rotate. The flow is axial at entry and exit. The fan is designed to produce a pressure difference, and hence force, to cause a flow through the fan. Factors which determine the performance of the fan include the number and shape of the blades. Fans have many applications including in wind tunnels and cooling towers. Design parameters include power, flow rate, pressure rise and efficiency.

The recharge oscillator model for El Niño–Southern Oscillation (ENSO) is a theory described for the first time in 1997 by Jin., which explains the periodical variation of the sea surface temperature (SST) and thermocline depth that occurs in the central equatorial Pacific Ocean. The physical mechanisms at the basis of this oscillation are periodical recharges and discharges of the zonal mean equatorial heat content, due to ocean-atmosphere interaction. Other theories have been proposed to model ENSO, such as the delayed oscillator, the western Pacific oscillator and the advective reflective oscillator. A unified and consistent model has been proposed by Wang in 2001, in which the recharge oscillator model is included as a particular case.

References

  1. Emanuel, K A. (1986). "An Air-Sea Interaction Theory for Tropical Cyclones. Part I: Steady-State Maintenance". Journal of the Atmospheric Sciences. 43 (6): 585–605. Bibcode:1986JAtS...43..585E. doi: 10.1175/1520-0469(1986)043<0585:AASITF>2.0.CO;2 .
  2. Bister, M.; Emanuel, K.A. (1998). "Dissipative heating and hurricane intensity". Meteorology and Atmospheric Physics. 65 (3–4): 233–240. Bibcode:1998MAP....65..233B. doi:10.1007/BF01030791. S2CID   123337988.
  3. Emanuel, K. (2000). "A Statistical Analysis of Tropical Cyclone Intensity". Monthly Weather Review. 128 (4): 1139–1152. Bibcode:2000MWRv..128.1139E. doi: 10.1175/1520-0493(2000)128<1139:ASAOTC>2.0.CO;2 .
  4. Knutson, T.R.; McBride, J.L.; Chan, J.; Emanuel, K.; Holland, G.; Landsea, C.; Held, I.; Kossin, J.P.; Srivastava, A.K.; Sugi, M. (2010). "Tropical cyclones and climate change". Nature Geoscience. 3 (3): 157–163. Bibcode:2010NatGe...3..157K. doi:10.1038/ngeo779. hdl: 11343/192963 .
  5. 1 2 Bister, M. (2002). "Low frequency variability of tropical cyclone potential intensity 1. Interannual to interdecadal variability". Journal of Geophysical Research. 107 (D24): 4801. Bibcode:2002JGRD..107.4801B. doi: 10.1029/2001JD000776 .
  6. Powell, M.D.; Vickery, P.J.; Reinhold, T.A. (2003). "Reduced drag coefficient for high wind speeds in tropical cyclones". Nature. 422 (6929): 279–83. Bibcode:2003Natur.422..279P. doi:10.1038/nature01481. PMID   12646913. S2CID   4424285.
  7. Bell, M.M.; Montgomery, M.T.; Emanuel, K.A. (2012). "Air–Sea Enthalpy and Momentum Exchange at Major Hurricane Wind Speeds Observed during CBLAST". Journal of the Atmospheric Sciences. 69 (11): 3197–3222. Bibcode:2012JAtS...69.3197B. doi:10.1175/JAS-D-11-0276.1. hdl: 1721.1/81202 . S2CID   17840178.
  8. Emanuel, K.; Sobel, A. (2013). "Response of tropical sea surface temperature, precipitation, and tropical cyclone-related variables to changes in global and local forcing". Journal of Advances in Modeling Earth Systems. 5 (2): 447–458. Bibcode:2013JAMES...5..447E. doi: 10.1002/jame.20032 . hdl: 1721.1/85564 .
  9. Woolnough, S. J.; Slingo, J. M.; Hoskins, B. J. (2000). "The Relationship between Convection and Sea Surface Temperature on Intraseasonal Timescales". Journal of Climate. 13 (12): 2086–2104. Bibcode:2000JCli...13.2086W. doi: 10.1175/1520-0442(2000)013<2086:TRBCAS>2.0.CO;2 .
  10. DeMaria, Mark; John Kaplan (September 1994). "Sea Surface Temperature and the Maximum Intensity of Atlantic Tropical Cyclones". Journal of Climate . American Meteorological Society. 7 (9): 1324–1334. Bibcode:1994JCli....7.1324D. doi: 10.1175/1520-0442(1994)007<1324:SSTATM>2.0.CO;2 . ISSN   1520-0442 . Retrieved 2008-07-30.