Skip to main content

Advertisement

Log in

Simulation of very severe cyclone Mala over Bay of Bengal with HWRF modeling system

  • Original Paper
  • Published:
Natural Hazards Aims and scope Submit manuscript

Abstract

Tropical cyclone is one of the most devastating weather phenomena all over the world. The Environmental Modeling Center (EMC) of the National Center for Environmental Prediction (NCEP) has developed a sophisticated mesoscale model known as Hurricane Weather Research and Forecasting (HWRF) system for tropical cyclone studies. The state-of-the-art HWRF model (atmospheric component) has been used in simulating most of the features our present study of a very severe tropical cyclone “Mala”, which developed on April 26 over the Bay of Bengal and crossed the Arakan coast of Myanmar on April 29, 2006. The initial and lateral boundary conditions are obtained from Global Forecast System (GFS) analysis and forecast fields of the NCEP, respectively. The performance of the model is evaluated with simulation of cyclone Mala with six different initial conditions at an interval of 12 h each from 00 UTC 25 April 2006 to 12 UTC 27 April 2006. The best result in terms of track and intensity forecast as obtained from different initial conditions is further investigated for large-scale fields and structure of the cyclone. For this purpose, a number of important predicted fields’ viz. central pressure/pressure drop, winds, precipitation, etc. are verified against observations/verification analysis. Also, some of the simulated diagnostic fields such as relative vorticity, pressure vertical velocity, heat fluxes, precipitation rate, and moisture convergences are investigated for understanding of the characteristics of the cyclone in more detail. The vector displacement errors in track forecasts are calculated with the estimated best track provided by the India Meteorological Department (IMD). The results indicate that the model is able to capture most of the features of cyclone Mala with reasonable accuracy.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11
Fig. 12

Similar content being viewed by others

References

  • Arakawa A, Lamb VR (1977) Computational design of the basic dynamical processes of the UCLA general circulation model. In: Chang J (ed) Methods in computational physics, vol 17. Academic Press, pp 173–265

  • Arakawa A, Schubert WH (1974) Interaction of a cumulus cloud ensemble with the large scale environment. Part I. J Atmos Sci 31:674–701

    Article  Google Scholar 

  • Bao J-W, Wilczak JM, Chio J-K, Kantha LH (2000) Numerical simulation of air–seas interaction under high wind conditions using a coupled model: a study of hurricane development. Mon Weather Rev 128:2190–2210

    Article  Google Scholar 

  • Black TL (1994) The new NMC mesoscale Eta model: description and forecast examples. Weather Forecast 9:265–278

    Article  Google Scholar 

  • Braun SA, Tao W-K (2000) Sensitivity of high-resolution simulations of hurricane Bob (1991) to planetary boundary layer parameterizations. Mon Weather Rev 128:3941–3961

    Google Scholar 

  • Dudhia J (1993) A non-hydrostatic version of Penn State–NCAR mesoscale model: validation tests and simulation of an Atlantic cyclone and cold front. Mon Weather Rev 121:1493–1513

    Article  Google Scholar 

  • Fels SB, Schwarzkopf MD (1975) The simplified exchange approximation: a new method for radiative transfer calculation. J Atmos Sci 32:1475–1488

    Article  Google Scholar 

  • Gopalakrishnan SG, Bacon DP, Ahmad NN, Boybeyi Z, Dunn TJ, Hall MS, Jin Y, Lee PCS, Madala RV, Ananthakrishna Sarma R, Turner MD, Wait T (2002) An operational multi-scale atmospheric model with grid adaptivity for hurricane forecasting. Mon Wea Rev 130(7):1830–1847

    Article  Google Scholar 

  • Grell GA (1993) Prognostic evaluation of assumptions used by cumulus parameterizations. Mon Wea Rev 121:764–787

    Article  Google Scholar 

  • Hong S–Y, Pan H–L (1996) Non-local boundary layer vertical diffusion in a medium-range forecast model. Mon Wea Rev 124:2322–2339

    Article  Google Scholar 

  • IMD (June 2007) Tracks of cyclones and depressions in the Bay of Bengal and Arabian Sea during 2006

  • Janjic ZI (1977) Pressure gradient force and advection scheme used for forecasting with steep and small scale topography. Contrib Atmos Phy 50:186–199

    Google Scholar 

  • Janjic ZI (1984) Non-linear advection schemes and energy cascade on semi-staggered grids. Mon Wea Rev 112:1234–1245

    Article  Google Scholar 

  • Janjic ZI (2003) A non-hydrostatic model based on a new approach. Met Atmos Phy 82:271–285

    Article  Google Scholar 

  • Janjic ZI, Gerrity JP Jr, Nickovic S (2001) An alternative approach to non-hydrostatic modeling. Mon Wea Rev 129:1164–1178

    Article  Google Scholar 

  • Lacis AA, Hansen JE (1974) A parameterization for the absorption of solar radiation in the earth’s atmosphere. J Atmos Sci 31:118–133

    Article  Google Scholar 

  • Liu Y, Zhang D–L, Yau MK (1997) A multi-scale numerical simulation of hurricane Andrew (1992). Part-I: explicit simulation and verification. Mon Wea Rev 125:3073–3093

    Article  Google Scholar 

  • Liu Y, Zhang D–L, Yau MK (1999) A multi-scale numerical simulation of hurricane Andrew (1992). Part-II: kinematics and inner core structure. Mon Wea Rev 127:2597–2616

    Article  Google Scholar 

  • Majumdar AB, Lele RR, Sunitha Devi S (2007) Cyclones and depressions over NIO during 2006. Mausam 58(3):303–322

    Google Scholar 

  • Mandal M, Mohanty UC, Sinha P, Ali MM (2007) Impact of sea surface temperature in modulating movement and intensity of tropical cyclones. Natural Hazard 41:413–427

    Article  Google Scholar 

  • Mesinger F (1977) Forward-backward scheme, and its use in a limited area model. Contrib Atmos Phys 50:200–210

    Google Scholar 

  • Mesinger F, Janjic ZI (1974) Noise due to time-dependent boundary conditions in limited area models. The GARP Programme on Numerical Experimentation, Rep No 4, WMO, Geneva, pp 31–32

  • Mohanty UC (1994) Tropical cyclones in the Bay of Bengal and deterministic methods for prediction of their trajectories. Sadhana 19(4):567–582

    Article  Google Scholar 

  • Mohanty UC, Gupta A (1997) Deterministic methods for prediction of tropical cyclone tracks. Mausam 48(2):257–272

    Google Scholar 

  • Mohanty UC, Mandal M, Raman S (2004) Simulation of Orissa super cyclone (1999) using PSU/NCAR mesoscale model. Nat Hazards 31(2):373–390

    Article  Google Scholar 

  • Mohanty UC, Pattanayak S, Osuri KK (2010) Changes in frequency and intensity of tropical cyclones over Indian seas in a warming environment. Disaster Dev 4:53–77

    Google Scholar 

  • Pattanayak S, Mohanty UC (2008) A comparative study on performance of MM5 and WRF models in simulation of tropical cyclones over Indian seas. Curr Sci 95(7):923–936

    Google Scholar 

  • Schwarzkopf MD, Fels SB (1985) Improvements to the algorithm for computing CO2 transmissivities and cooling rates. J Geophys Res 90:541–550

    Article  Google Scholar 

  • Schwarzkopf MD, Fels SB (1991) The simplified exchange method revisited: an accurate, rapid method for computations of infrared cooling rates and fluxes. J Geophys Res 96:9075–9096

    Article  Google Scholar 

  • Skamarock WC, Klemp JB, Dudhia J, Gill DO, Barker DM, Wang W, Powers JG (2005) A description of the advanced research WRF version 2

  • Wong KY, Yip CL (2005a) Tropical cyclone eye fix using genetic algorithm with temporal information. In: Proceedings of the 9th international conference on knowledge-based intelligent information and engineering systems (KES-2005). LNAI, Melbourne, Australia, pp 854–860

  • Wong KY, Yip CL (2005b) An intelligent tropical cyclone eye fix system using motion field analysis. In: Proceeding of the 17th IEEE international conference on tools with artificial intelligence (ICTAI-2005). Hong Kong, pp 652–656

  • Xue M, Droegemeier KK, Wong V (2000) The advanced regional prediction system (ARPS)—a multiscale nonhydrostatic atmospheric simulation and prediction tool. Part I: model dynamics and verification. Meteor Atmos Phys 75:161–193

    Article  Google Scholar 

  • Xue M, Droegemeier KK, Wong V, Shapiro A, Brewster K, Carr F, Weber D, Liu Y, Wang D-H (2001) The advanced regional prediction system (ARPS)—a multiscale nonhydrostatic atmospheric simulation and prediction tool. Part II: model physics and applications. Meteor Atmos Phys 76:134–165

    Article  Google Scholar 

  • Zhang D–L, Wang X (2003) Dependence of hurricane intensity and structure on vertical resolution and time-size. Adv Atmos Sci 20:711–725

    Article  Google Scholar 

Download references

Acknowledgments

The Indian National Center for Ocean Information Services (INCOIS), Ministry of Earth Sciences (MoES) is gratefully acknowledged for providing financial support to carry out this research. The authors acknowledge the Environmental Modeling Center (EMC) of NCEP for providing HWRF modeling system for the present study. The authors owe thanks to India Meteorological Department (IMD) for providing best track parameters of the storm and NASA for precipitation data sets for the validation of the model simulated results. The authors gratefully acknowledge the NCEP/NCAR for their analysis and forecast data sets used in the study.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Sujata Pattanayak.

Appendix 1

Appendix 1

1.1 HWRF modeling system

HWRF is a fully compressible Eulerian non-hydrostatic coordinate system with a hydrostatic option. The horizontal rotated latitude-longitude coordinate and the vertical terrain following hydrostatic-pressure-based sigma coordinate system is used. The sigma coordinate system is up to a specified pressure surface usually below the tropopause, and hydrostatic pressure coordinate above. Top of the model is a constant pressure surface.

1.2 Model formulation

The vertical sigma coordinate is defined as,

$$ \begin{aligned} \sigma =\, & \frac{{\pi - \pi_{t} }}{\mu } \\ \mu =\, & \pi_{s} - \pi_{t} \\ \end{aligned} $$

where π is the hydrostatic pressure, π s stands for the hydrostatic pressure at the surface, π t represents the hydrostatic pressure at the top of the model atmosphere.

Then the equations governing a dry, inviscid and adiabatic non-hydrostatic atmosphere are (Janjic et al. 2001),

$$ \frac{{\partial {\mathbf{v}}}}{\partial t} = - {\mathbf{v}} \cdot \nabla_{\sigma } {\mathbf{v}} - \dot{\sigma }\frac{{\partial {\mathbf{v}}}}{\partial \sigma } - (1 + \varepsilon )\nabla_{\sigma } \Upphi - \alpha \nabla_{\sigma } p + f{\mathbf{k}} \times {\mathbf{v}}\quad \left( {\text{Horizontal momentum}} \right) $$
$$ \frac{\partial T}{\partial t} = - {\mathbf{v}} \cdot \nabla_{\sigma } T - \dot{\sigma }\frac{\partial T}{\partial \sigma } + \frac{\alpha }{{C_{p} }}\left[ {\frac{\partial p}{\partial t} + {\mathbf{v}} \cdot \nabla_{\sigma } p + \dot{\sigma }\frac{\partial p}{\partial \sigma }} \right]\quad \left( {\text{Thermodynamic}} \right) $$
$$ \frac{\partial \mu }{\partial t} + \nabla_{\sigma } \cdot (\mu {\mathbf{v}}) + \frac{{\partial (\mu \dot{\sigma })}}{\partial \sigma } = 0\quad \left( {\text{Continuity}} \right) $$
$$ \frac{\partial p}{\partial \pi } = 1 + \varepsilon \quad \left( {\text{Vertical equation of motion}} \right) $$
$$ \frac{\partial \Upphi }{\partial \sigma } = - \mu \frac{RT}{p}\quad ({\text{Hypsometric}}) $$
$$ w = \frac{1}{g}\frac{{{\text{d}}\Upphi }}{{{\text{d}}t}} = \frac{1}{g}\left( {\frac{\partial \Upphi }{\partial t} + {\mathbf{v}} \cdot \nabla_{\sigma } \Upphi + \dot{\sigma }\frac{\partial \Upphi }{\partial \sigma }} \right)\quad ({\text{Non-hydrostatic continuity}}) $$
$$ \alpha = {\raise0.7ex\hbox{${RT}$} \!\mathord{\left/ {\vphantom {{RT} p}}\right.\kern-\nulldelimiterspace} \!\lower0.7ex\hbox{$p$}}\quad ({\text{Equation of state}}) $$
$$ \varepsilon = \frac{1}{g}\frac{{{\text{d}}w}}{{{\text{d}}t}}\quad ({\text{Definition of }}\varepsilon ) $$

where v is the horizontal wind vector, p is the non-hydrostatic pressure, R is the gas constant for dry air, T is temperature and Φ is geopotential.

It is to be noted that, Φ, w, and ε are not independent and there is no independent prognostic equations for w. Instead, w is computed diagnostically from the non-hydrostatic continuity equation. The parameter ε is the central point of attention for non-hydrostatic dynamics. In meso- and large-scale atmospheric flows, ε << 1.

1.3 Discretization of equations

The discretization of model equations is as follows.

1.4 Horizontal discretization equation

In order to organize the spurious accumulation of energy at the smallest resolvable scales, an energy and enstrophy conserving advection scheme for the semi-staggered E-grid is used (Arakawa and Lamb 1977).

The horizontal advection scheme for the E-grid is defined by,

$$ \begin{aligned} U = & \overline{\Updelta \pi }^{\lambda } u2\Updelta x \\ V = & \overline{\Updelta \pi }^{\varphi } v2\Updelta y \\ U^{\prime } = & \overline{\Updelta \pi }^{{\lambda^{\prime}}} \left( {\overline{u\Updelta y + v\Updelta x}^{\varphi \prime } } \right) \\ V^{\prime } = & \overline{\Updelta \pi }^{\varphi \prime } \left( {\overline{ - u\Updelta y + v\Updelta x}^{{\lambda^{\prime}}} } \right) \\ A = & 4\Updelta x\Updelta y,A^{\prime } = 2\Updelta x\Updelta y \\ \end{aligned} $$

where Δπ is the thickness of the layer in hydrostatic pressure and \( \Updelta x = a\cos (\varphi )\Updelta \lambda ,\Updelta y = a\Updelta \varphi . \) Then the horizontal advection for T, u and v are defined as,

$$ \begin{aligned} \frac{\partial T}{\partial t} = & - \frac{1}{\Updelta \pi }\left[ {\frac{1}{3}\frac{1}{A}\left( {\overline{{U\Updelta_{\lambda } T}}^{\lambda } + \overline{{V\Updelta_{\varphi } T}}^{\varphi } } \right) + \frac{2}{3}\frac{1}{{A^{\prime } }}\left( {\overline{{U^{\prime } \Updelta_{{\lambda^{\prime } }} T}}^{{\lambda^{\prime } }} + V\Updelta_{\varphi } T^{{\varphi^{\prime } }} } \right)} \right] + \cdots \\ \frac{\partial u}{\partial t} = & - \frac{1}{{\overline{\Updelta \pi }^{\lambda } }}\left[ {\frac{1}{3}\frac{1}{A}\left( {\overline{{\overline{U}^{\lambda } \Updelta_{\lambda } u}}^{\lambda } + \overline{{\overline{V}^{\lambda } \Updelta_{\varphi } u}}^{\varphi } } \right) + \frac{2}{3}\frac{1}{{A^{\prime } }}\left( {\overline{{\overline{{U^{\prime } }}^{\lambda } \Updelta_{{\lambda^{\prime } }} u}}^{{\lambda^{\prime } }} + \overline{{\overline{{V^{\prime } }}^{\lambda } \Updelta_{{\varphi^{\prime } }} u}}^{{\varphi^{\prime } }} } \right)} \right] + \cdots \\ \frac{\partial v}{\partial t} = & - \frac{1}{{\overline{\Updelta \pi }^{\varphi } }}\left[ {\frac{1}{3}\frac{1}{A}\left( {\overline{{\overline{U}^{\varphi } \Updelta_{\lambda } v}}^{\lambda } + \overline{{\overline{V}^{\varphi } \Updelta_{\varphi } v}}^{\varphi } } \right) + \frac{2}{3}\frac{1}{{A^{\prime } }}\left( {\overline{{\overline{{U^{\prime } }}^{\varphi } \Updelta_{{\lambda^{\prime } }} v}}^{{\lambda^{\prime } }} + \overline{{\overline{{V^{\prime } }}^{\varphi } \Updelta_{{\varphi^{\prime } }} v}}^{{\varphi^{\prime } }} } \right)} \right] + \cdots \\ \end{aligned} $$

In horizontal co-ordinate system, in order to achieve higher computational efficiency of the model code, a transformed latitude-longitude coordinate system is used. This system is obtained by rotation of the natural geodesic latitude-longitude in such a way that the intersection of the equator and zero meridian of the transformed system (λ 0, φ0) coincides with the center of the model domain. The transformation and inverse transformation equations between the rotated and the natural latitude-longitude system (λ, φ) are as follows .

$$ \begin{aligned} \Uplambda = & \arctan \frac{{\cos \varphi \sin (\lambda - \lambda_{0} )}}{{\cos \varphi_{0} \cos \varphi \cos (\lambda - \lambda_{0} ) + \sin \varphi_{0} \sin \varphi }} \\ \Upphi = & \arcsin (\cos \varphi_{0} \sin \varphi - \sin \varphi_{0} \cos \varphi \cos (\lambda - \lambda_{0} )) \\ \varphi = & \arcsin (\sin \varphi_{0} \cos \Upphi \cos \Uplambda + \cos \varphi_{0} \sin \Upphi \\ \lambda =\, & \lambda_{0} + \arcsin \left( {\frac{\sin \Uplambda \cos \Upphi }{{\cos \varphi_{0} }}} \right) \\ \end{aligned} $$

The horizontal wind in the rotated system V = (U, V) expressed in terms of the wind in the natural latitude/longitude system v = (u,v) is defined as;

$$ \begin{aligned} U = & \frac{{u\left[ {\cos \varphi_{0} \cos \varphi + \sin \varphi_{0} \sin \varphi \cos (\lambda - \lambda_{0} )} \right] - v\sin \varphi_{0} \sin (\lambda - \lambda_{0} )}}{{\sqrt {1 - \left[ {\cos \varphi_{0} \sin \varphi - \sin \varphi_{0} \cos \varphi \cos (\lambda - \lambda_{0} )} \right]^{2} } }} \\ V = & \frac{{u\sin \varphi_{0} \sin (\lambda - \lambda_{0} ) + v\left[ {\cos \varphi_{0} \cos \varphi + \sin \varphi_{0} \sin \varphi \cos (\lambda - \lambda_{0} )} \right]}}{{\sqrt {1 - \left[ {\cos \varphi_{0} \sin \varphi - \sin \varphi_{0} \cos \varphi \cos (\lambda - \lambda_{0} )} \right]} }} \\ \end{aligned} $$

The inverse wind transformation is given by,

$$ \begin{aligned} u = & \frac{{U(\cos \varphi_{0} \cos \Upphi - \sin \varphi_{0} \sin \Upphi \cos \Uplambda ) + V\sin \varphi_{0} \sin \Uplambda }}{{\sqrt {1 - (\cos \varphi_{0} \sin \Upphi + \sin \varphi_{0} \cos \Uplambda \cos \Upphi )^{2} } }} \\ v = & \frac{{ - U\sin \varphi_{0} \sin \Uplambda + V(\cos \varphi_{0} \cos \Upphi - \sin \varphi_{0} \sin \Upphi \cos \Uplambda }}{{\sqrt {1 - (\cos \varphi_{0} \sin \Upphi + \sin \varphi_{0} \cos \Uplambda \cos \Upphi )^{2} } }} \\ \end{aligned} $$

1.5 Vertical discretization equations

The Lorenz vertical staggering of the variables is used in the vertical (Janjic 1977). The geopotential and non-hydrostatic pressure are defined at the interfaces of the layers, while all three velocity components and temperature are carried in the middle layers of the model. The vertical velocity is defined at the E-grid mass points.

The terrain following hydrostatic-pressure-based sigma coordinate up to a specified pressure surface usually below the tropopause, and hydrostatic pressure coordinate above. Top of the model is constant pressure surface.

In the pressure range,

$$ \nabla_{p} \cdot ({\mathbf{v}}) + \frac{\partial \omega }{\partial p} = 0 $$

In the sigma range,

$$ \frac{\partial PD}{\partial t} + \nabla_{\sigma } \cdot (PD\,{\mathbf{v}}) + \frac{{\partial (PD\dot{\sigma })}}{\partial \sigma } = 0 $$

where as \( PD\dot{\sigma } = \omega \) at the vertical level of the model where sigma range changes to the pressure range.

Hence, the mass continuity equation can be written as:

For the pressure range:

$$ 0 = - \left[ {\frac{1}{3}\frac{1}{A}\left( {\Updelta_{\lambda } U + \Updelta {}_{\varphi }V} \right) + \frac{2}{3}\frac{1}{{A^{\prime } }}\left( {\Updelta_{{\lambda^{\prime } }} U^{\prime } + \Updelta_{{\varphi^{\prime } }} V^{\prime } } \right)} \right] - \Updelta_{\pi } \omega $$

For the sigma range:

$$ \frac{\partial \Updelta \pi }{\partial t} = - \left[ {\frac{1}{3}\frac{1}{A}\left( {\Updelta_{\lambda } U + \Updelta {}_{\varphi }V} \right) + \frac{2}{3}\frac{1}{{A^{\prime } }}\left( {\Updelta_{{\lambda^{\prime } }} U^{\prime } + \Updelta_{{\varphi^{\prime } }} V^{\prime } } \right)} \right] - \Updelta_{\sigma } \left( {\mu \dot{\sigma }} \right) $$

For the interface between sigma and pressure range:

$$ (\mu \dot{\sigma })_{I} = \omega_{I} $$

1.6 Time discretization equations

HWRF uses four types of time integration: (i) modified Adams–Bashforth scheme for horizontal advection of u, v and T, and Coriolis terms, (ii) Crank–Nicholson for vertical advection of u, v and T, (iii) forward–backward scheme for adjustment terms and (iv) an implicit scheme for vertically propagating sound waves.

The Adam-Bashforth scheme can be represented as:

$$ \frac{{y^{\tau + 1} - y^{\tau } }}{\Updelta t} = \frac{3}{2}f\left( {y^{\tau } } \right) - \frac{1}{2}f\left( {y^{\tau - 1} } \right) $$

This method has a slight linear instability, which can be stabilized as;

$$ \frac{{y^{\tau + 1} - y^{\tau } }}{\Updelta t} = 1.533f\left( {y^{\tau } } \right) - 0.533f\left( {y^{\tau - 1} } \right) $$

The Crank–Nicholson scheme can be represented as:

$$ \frac{{y^{\tau + 1} - y^{\tau } }}{\Updelta t} = \frac{1}{2}\left[ {f\left( {y^{\tau + 1} } \right) + f\left( {y^{\tau } } \right)} \right] $$

This is an implicit scheme and so is always stable.

The time integration for the terms involving the propagation of gravity waves is handled by a forward–backward process.

Using the shallow water equation,

$$ \frac{\partial u}{\partial t} = - g\frac{\partial h}{\partial x},\quad \frac{\partial h}{\partial t} = - H\frac{\partial u}{\partial x} $$

The mass tendency equation is advanced by a forward step:

$$ h^{\tau + 1} = h^{\tau } - \Updelta t\;H\;\frac{{\partial u^{\tau } }}{\partial x} $$

The velocity equations are then advanced with a backward step:

$$ u^{\tau + 1} = u^{\tau } - \Updelta t\;g\;\frac{{\partial h^{\tau + 1} }}{\partial x} $$

The time integration for vertically propagating sound waves is:

$$ \frac{{\partial^{2} p^{\prime } }}{{\partial t^{2} }} = \frac{{p^{\prime n + 1} - 2p^{\prime n} + p^{\prime n - 1} }}{{\Updelta t^{2} }} = \frac{{{\mathbf{c}}_{p} }}{{{\mathbf{c}}_{v} }}R\,T_{0} \;\frac{{\partial^{2} p^{\prime n + 1} }}{{\partial z_{0}^{2} }}, $$

where p′ represents a perturbation pressure

The advection of other variables such as water vapor and liquid water uses the Langangian forward time differencing, which can be represented as:

$$ \frac{{y_{j}^{\tau + 1 * } - y_{j}^{\tau } }}{\Updelta t} = - u\frac{{y_{j}^{\tau } - y_{j - 1}^{\tau } }}{\Updelta x},\quad u > 0 $$

This is followed by a negative diffusion step to reduce smoothing:

$$ \frac{{y_{j}^{\tau + 1} - y_{j}^{\tau + 1 * } }}{\Updelta t} = - f\left( {\frac{u\Updelta t}{\Updelta x}} \right)\frac{{y_{j - 1}^{\tau + 1 * } - 2y_{j}^{\tau + 1 * } + y_{j + 1}^{\tau + 1 * } }}{\Updelta x},\quad f\left( {\frac{u\Updelta t}{\Updelta x}} \right) > 0 $$

With the non-hydrostatic approach, with vanishing ɛ, the time discrete equation with adding Crank–Nicholson vertical and Adams–Bashforth horizontal advection can be written as:

$$ T_{1}^{ * } = T^{n} - \Updelta t\left( {{\mathbf{v}}^{n} \cdot \nabla_{\sigma } T^{n + 1/2} + \dot{\sigma }^{n} \frac{{\partial T^{n + 1/2} }}{\partial \sigma }} \right) $$

Finally

$$ T_{1} = T_{1}^{ * } + \frac{\Updelta t}{{{\mathbf{c}}_{p} }}\,\frac{{RT^{n} }}{{p^{n} }}\omega_{1} $$

1.7 Lateral boundary condition

The velocity and mass variables are specified only on the outermost rows and columns. On the first row or column from the outer boundaries, each variable is replaced by a four point average from its surrounding points (Mesinger and Janjic 1974, Mesinger 1977). Inside the domain, the initial values before the averaging are defined by the tendency equations with the advection terms computed using upstream differences on the three rows next to the internal boundary where the four-point averaging is applied. The enhanced divergence damping is generally applied close to the boundaries.

The vertical boundary conditions are:

$$ \dot{\sigma } = 0\quad {\text{and}}\quad p - \pi = 0\;{\text{at}}\;\sigma = 0 $$

and

$$ \dot{\sigma } = 0\;\quad {\text{and}}\quad \;\frac{\partial (p - \pi )}{\partial \sigma } = 0\;{\text{at}}\;\sigma = 1 $$

Rights and permissions

Reprints and permissions

About this article

Cite this article

Pattanayak, S., Mohanty, U.C. & Gopalakrishnan, S.G. Simulation of very severe cyclone Mala over Bay of Bengal with HWRF modeling system. Nat Hazards 63, 1413–1437 (2012). https://doi.org/10.1007/s11069-011-9863-z

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11069-011-9863-z

Keywords

Navigation