1. Introduction
In recent years, the nonhydrostatic models have been used for predictions of climate change with long time integrations. Directly calculating interactions of water vapors, clouds, and radiation for several tens of days, radiative–convective equilibriums are obtained in a large domain with horizontal length from 100 to 1000 km (Held et al. 1993; Tompkins and Craig 1998b, 1999). Some research groups, on the other hand, try to extend nonhydrostatic models to cover the entire globe (Semazzi et al. 1995; Cullen et al. 1997; Qian et al. 1998; Côté et al. 1998a,b; Smolarkiewicz et al. 1999). In the near future, development of computer facilities will allow us to use high-resolution global climate models with horizontal resolution around 5–10 km. In those days, we will need to choose the nonhydrostatic equations as the dynamical framework of the climate models.
The nonhydrostatic models, or the cumulus resolving models, are nowadays developed and used by many research groups, and have shown successful simulations of mesoscale convections. These models are, however, originally intended to be used for short-range integrations of, say, a few days, so that much attention is not being paid to conservation of physical quantities. Although one needs to integrate for several tens of days to obtain a radiative–convective equilibrium (Tompkins and Craig 1998a), the conservations of mass and energy were not fully discussed in the nonhydrostatic model studies for such long-range integrations. One may think that the conservations of mass and energy are not the primary requirement for calculations of radiative–convective equilibrium or climate modeling, since there remains uncertainty derived from the artificial numerical smoothing even when the domain integral quantities are conserved. In the case that the numerical scheme does not guarantee the conservations, however, reliability of model results would be reduced at least to the range of fluctuations of the conserved quantities.
The nonhydrostatic equations have less approximations to the governing equations of the numerical models of the atmosphere, in comparison to the hydrostatic equations usually used for the large-scale models. The nonhydrostatic equations system is categorized into two groups: the incompressible system (Ogura and Phillips 1962) and the compressible system. Since the density is not a variable quantity in the incompressible system, we only consider the compressible nonhydrostatic equations in this paper. Essentially, the compressible nonhydrostatic equations are equivalent to the Euler equations in the fluid dynamics since they do not introduce any approximations to the governing equations of a fluid. As in the case of the Euler equations, the equations are formulated with the conservative forms of natural variables, that is, density, momentum, and total energy. Therefore, if these variables are discretized in the flux forms, we will obtain a conservative numerical scheme of the nonhydrostatic equations.
Such an approach according to fluid dynamics has not usually been taken in the nonhydrostatic modeling to simulate mesoscale convections. Instead of density and total energy, pressure p (or the Exner function π) and potential temperature θ (or temperature T) are usually used as prognostic variables. Historically, the choice of these variables may be due to the fact that p and T are directly measured from observations. Furthermore, The quasi-Boussinesq approximation of a deep atmosphere, which is the base of the incompressible nonhydrostatic equations, is formulated by using π and θ by Ogura and Phillips (1962). It is also an advantage of using potential temperature since θ is conserved in the Lagrangian sense in adiabatic motions and can be used as a tracer in simulations of the mesoscale convection.
The recent advance of computing resources has invoked the need for climate modeling with the nonhydrostatic equations. The importance of conservations of mass and energy is being pointed out and becoming appreciated. Doms and Schättler (1997) suggest a transition to use the equations of using density and internal energy as prognostic variables instead of pressure and enthalpy. Taylor (1984) and Gallus and Rančić (1996) have proposed the energy conserving schemes by using enthalpy or temperature as a prognostic variable. Recently, Klemp et al. (2000) have devised a conservative form of discretized equations by using density as a prognostic variable in addition to a flux form potential temperature. Xue et al. (2000) have also proposed a conservative form of discretized equations with minimum approximations to the original governing equations.
On the other hand, most of the currently used global models in the hydrostatic equations are based on the conservative forms. Arakawa and Lamb (1977) and Arakawa and Suarez (1983) developed an energy conserving scheme by considering a transformation of the kinetic energy to enthalpy based on the pressure coordinates. As an extension of this approach, there is a category of the nonhydrostatic models in which pressure in the hydrostatic balance is used as a vertical coordinate (Laprise 1992; Juang 1992; Gallus and Rančić 1996; Klemp et al. 2000). In particular, Gallus and Rančić (1996) took care of the conservation of energy in their nonhydrostatic model with the pressure coordinate. It is true that the use of the pressure coordinate in nonhydrostatic models is advantageous; it is easy to incorporate observational data to the models and the model results can be directly compared with those of the large-scale models. This does not mean, however, that the use of the pressure coordinate is the only way to the energy conserving scheme.
We think that as a horizontally resolvable scale of models becomes smaller, such as below 10 km, the geometric height z coordinate is a more appropriate choice for the vertical coordinate, since the assumption of the hydrostatic balance becomes less relevant. The height-based coordinate is used in many groups of the nonhydrostatic models, and will be an important candidate when the nonhydrostatic models are extended to the global model. In this respect, we develop a conservative numerical scheme using the height coordinate in the compressible nonhydrostatic equations. Our approach is similar to that of Taylor (1984). Although Taylor (1984) reported the vertical discretization method for the conservation of energy and entropy, we particularly concentrate on the time discretization in order to stably calculate propagation of sound waves. We also demonstrate some numerical experiments to show the usefulness of the scheme, while Taylor (1984) did not show any application of his scheme.
The structure of this paper is as follows. In section 2, we formulate a conservative scheme of the nonhydrostatic model that will be used in the following section. In particular, in section 3, the conservation of energy is discussed and alternative forms of the energy conservation are argued. In section 4, the numerical scheme is applied to some test simulations. Section 5 summarizes the results and compares them with other research.
2. Numerical scheme
a. Objectives
We summarize first the requirements of the new scheme to be developed. We are in the course of developing a global climate model in the nonhydrostatic equations to operate with a ultramassive parallel computer with distributed memories. The proposed new scheme is for the purpose of this target. First, we require conservations of variables; in particular, conservations of mass and energy are important for the climate modeling. To satisfy the conservations, we discretized the equations in the flux forms with the finite volume method. Second, the use of a massive parallel computer gives a restriction on the choice of the time integration scheme. We are only considering a horizontally two-dimensional decomposition of the domain for the parallel computation, that is, the whole horizontal computational domain consists of tiles of a rectangular area, each of which is assigned to one of computer nodes. In order to efficiently run on massive parallel computers, data communications between nodes should be smaller. Particularly, we should avoid those methods that generate global communications among many nodes, or large transfer of memory between nodes. The spectrum method and the finite element method are counted as such methods. Use of the horizontally implicit method for sound and gravity waves or the anelastic equations should be reserved since they require elliptic equations to be solved. Solving an elliptic equation on massive parallel computers usually generates large communications between nodes, although efficient numerical methods such as a multigrid method are nowadays available.
Choice of the time integration scheme greatly depends on the treatment of sound waves. We choose the horizontally explicit and vertically implicit scheme for sound waves based on the reason described above. For a horizontally explicit scheme, one also can use the time splitting scheme (Klemp and Wilhelmson 1978; Skamarock and Klemp 1992) to accelerate the efficiency; only the terms related to sound waves are integrated by a small time step and the other slower terms such as advection terms are integrated by a large time step. The scheme shown below can incorporate the time splitting scheme, although the results with the time splitting will be omitted throughout the paper, since the conservation properties are the main concern of the present study.
We start from the Euler equations and make the least approximations to them. The Euler equations are written in the conservative forms of density, momentum, and total energy. The equation of total energy is converted to that of the internal energy, which has transformation terms with the kinetic and potential energies. We choose density ρ and internal energy e as prognostic variables. This choice is contrasted with that in the usual nonhydrostatic models where pressure p (or the Exner function π) and potential temperature θ (or temperature T) are used as prognostic variables. Here, the Exner function is defined as π =
b. The basic equations












c. The pressure equation






The following problems arise, however, if the pressure equation is used as a prognostic equation. First, the conservation of mass is not exactly satisfied. In fact, in the case that cs is not constant, a volume integral of p′ does not conserve when Eq. (14) is discretized in the flux form. Even when cs is constant, it is not the total mass but the domain integral of pressure that is conserved.






In the case of no gravity, the contribution from the right-hand side of Eq. (17) vanishes in the linearized equation around the basic reference state. The convergence of the advective flux
d. Time discretization


























In the above formula, we derived the Helmholtz equation for Wn+1. The other choice of the equation for Rn+1 or Pn+1 may be considered. One of the advantages of solving for W is that the top and the bottom boundary conditions are easily incorporated; in the case when the kinematic condition is applied at the boundaries, the boundary values of Wn+1 are directly given. In addition, round-off errors will be smaller if Rn+1 and En+1 are directly integrated using the flux form equations. In particular, by using the energy equation (47), we can make a correction for the errors introduced when Eq. (36) is derived. Furthermore, if the specific heat depends on temperature or humidity, Eq. (36) is only an approximation. Equation (47) satisfies the conservation since the advection term is written in the flux form.
e. Terrain-following coordinate




















3. Energy budget
a. Transformation of energy
In section 2d, we described the time discretization of the model equations. We have pointed out that each of the transformation terms of energy should be consistently discretized to satisfy the conservation of energy, and have presented one possible approach. In this section, we consider how the conservation of energy is satisfied in the numerical scheme and argue about alternative forms of the equation of energy.














It is clear from Eq. (71) that the conservation of kinetic energy is not exactly satisfied, since
b. Integration of the sum of internal energy and kinetic energy
As a totally different approach to the exact conservation of total energy, one could choose the sum of internal energy and kinetic energy as a prognostic variable. We have implicitly solved for the variables R, W, and P from Eqs. (42), (43), and (44) for the vertical propagation of sound waves. In general, the internal energy E obtained by the following method does not satisfy the relation (36) with P. Nevertheless, as in the same thought of using Eq. (47) for the prediction of E, the following method may be used as an iterative process, if the difference of E obtained by the two methods is small.




c. Integration of entropy


The similar approach is adopted by Taylor (1984) who followed Arakawa and Lamb (1977) and devised the vertical discretization scheme using the height z as the vertical coordinate. Taylor uses the equation of enthalpy as a prognostic equation, and guarantees the conservations of potential temperature and entropy by appropriate vertical interpolations of thermodynamic variables. If we use such a vertical discretization as Taylor's, potential temperature and entropy will be also conserved even when the internal energy is chosen as a prognostic variable.
4. Model and results
a. Model
We show some results for simple simulations using the numerical scheme described in sections 2 and 3. We have proposed three methods of the energy integration. The first method is based on Eq. (44), from which En+1 is given by multiplication of Cυ/Rd to Pn+1: this is referred to as the “noncorrection” method. The second is based on Eq. (47) where the correction is made to the transformation terms of energies: this method is referred to as the “correction” method. In the third method, the conservation of the total energy is guaranteed based on Eq. (74): this will be called the “conservative” method. The results of using these methods are compared in some of the following experiments.
The model is developed in the three-dimensional Cartesian coordinate, but only the results for the two-dimensional calculations are shown here. The definition points of the variables are based on the Arakawa C grid; R, E, and P are defined at the same points, and U, V, and W are defined at the staggered points shifted half point in the x, y, and z directions, respectively. The Lorenz grid is used in the vertical direction. Rigid walls are placed at the bottom and the top of boundaries and free slip and insulating conditions are used at the boundaries. Periodicity is assumed at the lateral boundaries. The time integration of sound waves is based on the forward–backward scheme; the horizontally explicit and vertically implicit scheme as described in section 2d. The model is constructed following the time splitting method (Klemp and Wilhelmson 1978; Skamarock and Klemp 1992), but we simply set the small time step equal to the large time step. The large time step is integrated with the leapfrog scheme with the time filter to suppress the computational mode. We denote the time step of the leapfrog scheme as 2Δt, where 2Δt is equal to the time step of the forward–backward scheme of sound waves. We use a simple second-order centered-in-space scheme for the advections.
b. One-dimensional vertical propagation of sound waves
First, we calculate vertically propagating sound waves to see the stability of the scheme. A perturbation of pressure is given in a layer at the initial state under a horizontally uniform condition. The initial state is uniform temperature 250 K at rest in the hydrostatic balance except for the perturbation of pressure p′ = 100 hPa between the heights 2.5 and 5 km. The top of the atmosphere is zT = 15 km, and the layer difference is uniform with Δz = 500 m. No numerical smoothing such as numerical diffusion, Rayleigh damping, or the time filter is introduced in this case. Note that, however, the fully implicit scheme has a damping characteristic (see the appendix).
First, we use the correction method for the energy integration. Figures 1 and 2 show the vertical distributions of pressure and vertical velocity, respectively, at t = 10, 20, and 30 s. We compare the profiles with the time step Δt = 0.1, 1, and 10 s. It can be confirmed that the case with Δt = 10 s, which is larger than the Courant–Friedrichs–Lewy (CFL) criterion for the sound wave, is stably integrated. Since the fully implicit scheme is used for the vertical propagation of sound waves, the amplitude of pressure perturbation rapidly decreases as Δt becomes larger (see the appendix). The characteristics of these figures are almost the same for the three methods of the energy integration. The difference between the methods is shown if the energy budget is examined.


c. Horizontal propagation of gravity waves


Figure 4 shows the distributions of potential temperature at the initial state and t = 3000 s. Although a slight asymmetry emerges in the vertical direction, the horizontal propagation of the gravity waves almost agrees with the analytic solution shown by Skamarock and Klemp (1994).
Figure 5 shows time sequences of the volume-averaged values of energies. The differences from the initial values are shown. Similar to Fig. 3, the conservation of total energy is well satisfied. In particular, as shown by Fig. 5b, the transformation between the kinetic energy and the available energy is appropriately calculated. The change in mass is negligibly small though the mass is increasing due to the round-off error: the change in the mean density is below 10−13 kg m−3.
d. Topographic waves


Figure 6 shows the structure of the perturbation of horizontal velocity u′ and the vertical velocity w at t = 3000 s. The amplitudes of the velocities are about 1000 times larger than those of the analytic solution shown by Durran and Klemp (1983, their Figs. 1 and 2). At this time t = 3000 s (
Figure 7 shows time sequences of energies. Since the damping terms are included above ξ = 8 km, the change in the total energy should be equal to the accumulation of the energy sinks due to the damping terms. In Fig. 7b, the accumulated energy change due to the damping terms is shown by the dashed–dotted curve ER. It is found that ER is almost explained by the damping term due to temperature perturbations (not shown). Figure 7b shows the difference between the change in the total energy and ER by the dotted curve. It is confirmed that the change in the total energy is almost equal to ER. Since this simulation is essentially linear, the energy conservation is well satisfied if the correction method is used. The mass is also almost conserved; the change in the mean density is about 10−14 kg m−3 for the 3000-s integration. This comes from the round-off errors.
e. Cold-bubble simulation


The distributions of the perturbation of potential temperature at t = 0, 300, 600, and 900 s are shown in Fig. 8 for the 50-m grid simulation. The gravity current is propagating along the lower boundary and three rotors associated with the Kelvin–Helmholtz shear instability are well captured. The distribution of potential temperature at t = 900 s is comparable to the corresponding resolution experiment shown by Straka et al. (1993). Figure 9 compares the distributions of the perturbation of potential temperature at t = 900 s in the cases with the coarser resolutions 100 and 200 m. Since we only use the second-order centered-in-space scheme for the advection, the internal structure of vortices are not well simulated particularly in the case with the 200-m grid. However, as the resolution becomes finer, the representation of the rotors becomes better; this results indicate the confidence of the present scheme. The maximum and minimum values of perturbation of potential temperature at t = 900 s is 0.007000 K, −12.1084 K for the 50-m grid, 0.2178 K, −16.9435 K for the 100-m grid, and 2.2038 K, −21.6496 K for the 200-m grid. These values tend to converge to those of the reference solution presented by Straka et al. (1993).
Time sequences of energies for the simulation with the 50-m grid are displayed in Fig. 10. The nonlinear advective terms play important roles in this simulation. Strictly, the conservation of total energy is not satisfied in the correction scheme, if the advection of the kinetic energy is important. Reflecting this fact, the total energy is decreased by 5.12 J m−3 for the 900-s calculation as shown by Fig. 10b.
Gallus and Rančić (1996) developed a nonhydrostatic model with an energy conserving scheme using the pressure in the hydrostatic balance as a vertical coordinate. Gallus and Rančić (1996) have considered only the transformation term due to the pressure gradient force between the internal energy and the kinetic energy. Their scheme does not guarantee the exact conservation of total energy, hence it has the same limitation with the correction method. They also have shown the result of the cold-bubble experiment by using a grid interval of 100 m. In their calculation, their total energy is defined as the sum of the internal energy and the kinetic energy [Eq. (A.5) of Gallus and Rančić]. They stated that the change in the total energy is about 0.05% for the 40 000 iterations in the 900-s integration. In our calculation, since the initial value of the internal energy is 175 868 J m−3, the corresponding change in the energy is 0.003% in the 18 000 steps integration. It may be said our result is superior to that of Gallus and Rančić in terms of the conservation of total energy.
As an alternative approach, we have performed the similar calculation of the cold-bubble experiment based on the conservative method. We found that the simulation is successfully performed without any instability. Figure 10c shows time sequences of energies, which are compared with Fig. 10b. The change in the total energy during 900 s is only −0.005 428 J m−3 for the 50-m grid, −0.010 84 J m−3 for the 100-m grid, and −0.025 28 J m−3 for the 200-m grid; the conservation is much improved. This result suggests that the use of the conservation method is very encouraging.
5. Summary and discussions
We have devised a new dynamical scheme with the conservative forms of the nonhydrostatic models. It is aimed to apply to a global climate model for long time integrations in the future. In contrast to the past models based on the pressure and potential temperature as prognostic variables, the density and the internal energy are integrated in the flux forms. As a result, the conservation of total mass is satisfied within round-off errors. The conservation of total energy is approximately satisfied by considering the discretized forms of the major transformation terms of energy, though the exact conservation is not guaranteed since the inconsistency in the transformation of the nonlinear terms of the kinetic energy. We also argue about an alternative method in which the sum of internal energy and kinetic energy is used as a prognostic variable and the exact conservation of total energy is satisfied. We incorporated these schemes into a nonhydrostatic model with the horizontally explicit and vertically implicit time integration scheme for sound waves and performed various numerical experiments for the dry atmosphere. The numerical results show that the performance is comparative to that of the past studies, and that the proposed scheme is promising. In particular, the conservation of mass is confirmed with a great accuracy. The conservation of total energy is also demonstrated for the vertical propagation of sound waves, the horizontal propagation of gravity waves, and the finite amplitude topographic waves. The density current induced by an initial cold bubble is also calculated, which shows the approximate conservation of total energy. The propagation of the density current is successfully simulated if the sum of internal energy and kinetic energy is used as a prognostic variable. In this case, the conservation of total energy is much improved.
As another difference, they use ρw* instead of W = ρw for the implicit calculation of the terrain following coordinate [Eq. (58)]. As shown by Eq. (58), the horizontal velocity components u and υ are needed to calculate w*. We solve for ρw but not for ρw* in the implicit calculation; the vertical velocity is given by w = Jxu + Jyυ at the bottom boundary. In the integration of density, the convergence term associated with the vertical velocity is divided into two terms, one with Jxu + Jyυ and the other with w; this division might introduce relatively larger round-off errors. But the errors are small, and we think the choice of ρw is not a critical for the model performance.
When the time splitting method is introduced in the time integration, Klemp et al. (2000) propose that the advection term of energy should be split into the slow mode and the fast mode. They argue that the advection in the small time step is the deviation of the total advection from the advection related to the slow mode, and that sophisticated advection schemes such as semi-Lagrangian schemes should be incorporated into the discretization of the slow mode. Although we did not use any elaborated advection scheme and did not split the fast mode and the slow mode in the present simulations, introduction of the splitting of the advection term is worth considering for efficient calculations. We have done some preliminary tests of the time splitting for various experiments including horizontally propagating sound waves with basic winds. We have confirmed that the numerical calculation is stable even when the large time step is 10 times larger than the small time step.
In the present study, we only performed the simulations in the dry atmosphere. The properties of the conservations should be further examined in the case with the important hydrodynamic processes related to latent heat release. We will calculate the radiative–convective equilibrium in the two- or three-dimensional domains for the long time integrations such as several days (e.g., Held et al. 1993; Tompkins and Craig 1998b, 1999) and examine the conservations of the mass and the energy. We are planning to extend this dynamical model framework to the sphere and develop a global nonhydrostatic model for the climate modeling.
Acknowledgments
The author is grateful to Y. Kurihara, H. Tomita, M. Tsugawa, and T. Nasuno for helpful comments. The numerical calculations of the present study are done using HITACH SR8000 at the University of Tokyo under cooperative research with the Center for Climate System Research, University of Tokyo, NEC SX4 at the National Institute of Environmental Studies, NEC SX5 at the Frontier Research System for Global Change, and the parallel computers at the High-Tech Research Center at Saitama Institute of Technology.
REFERENCES
Arakawa, A., and V. R. Lamb, 1977: Computational design of the basic dynamical processes of the UCLA general circulation model. Methods Comput. Phys., 17 , 173–265.
Arakawa, A., and M. J. Suarez, 1983: Vertical differencing of the primitive equations in sigma coordinates. Mon. Wea. Rev., 111 , 34–45.
Côté, J., S. Gravel, A. M'ethot, and A. Patoine, 1998a: The operational CMC–MRB global environmental multiscale (GEM) model. Part I: Design considerations and formulation. Mon. Wea. Rev., 126 , 1373–1395.
Côté, J., J-G. Desmarais, S. Gravel, A. Méthot, A. Patoine, M. Roch, and A. Staniforth, 1998b: The operational CMC–MRB global environmental multiscale (GEM) model. Part II: Results. Mon. Wea. Rev., 126 , 1397–1418.
Cullen, M. J. P., T. Davies, M. H. Mawson, J. A. James, S. C. Coutler, and A. Malcolm, 1997: An overview of numerical methods for the next generation U.K. NWP and climate model. Numerical Methods in Atmospheric and Oceanic Modelling, The Andrew J. Robert Memorial Volume, C. A. Lin et al. Eds., NRC Research Press, 425–444.
Doms, G., and U. Schättler, 1997: The nonhydrostatic limited-are a model LM (Lokal–Modell) of DWD. Part I: Scientific Documentation. Deutscher Wetterdienst, 155 pp.
Durran, D. R., and J. Klemp, 1983: A compressible model for the simulation of moist mountain waves. Mon. Wea. Rev., 111 , 2341–2361.
Gallus, W. Jr,, and M. Rančić, 1996: A non-hydrostatic version of the NMC's regional Eta model. Quart. J. Roy. Meteor. Soc., 122 , 495–513.
Held, I. M., R. S. Hemler, and V. Ramaswamy, 1993: Radiative–convective equilibrium with explicit two-dimensional moist convection. J. Atmos. Sci., 50 , 3909–3927.
Juang, H-M. H., 1992: A spectral fully compressible nonhydrostatic mesoscale model in hydrostatic sigma coordinates: Formulation and preliminary results. Meteor. Atmos. Phys., 50 , 75–88.
Klemp, J. B., and R. B. Wilhemson, 1978: The simulation of three-dimensional convective storm dynamics. J. Atmos. Sci., 35 , 1070–1096.
Klemp, J. B., W. C. Skamarock, and J. Dudhia, cited 2000: Conservative split-explicit time integration methods for the compressible nonhydrostatic equations. [available online at http://wrf-model.org/WG1/wg1_main.html.].
Laprise, R., 1992: The Euler equations of motion with hydrostatic pressure as independent variable. Mon. Wea. Rev., 120 , 197–207.
Morinishi, Y., T. S. Lund, O. V. Vasilyev, and P. Moin, 1998: Fully conservative higher order finite difference schemes for incompressible flow. J. Comput. Phys., 143 , 90–124.
Ogura, Y., and A. Phillips, 1962: Scale analysis of deep and shallow convection in the atmosphere. J. Atmos. Sci., 19 , 173–179.
Ooyama, K. V., 1990: A thermodynamic foundation for modeling the moist atmosphere. J. Atmos. Sci., 47 , 2580–2593.
Qian, J-H., F. H. M. Semazzi, and J. S. Scroggs, 1998: A global nonhydrostatic semi-Lagrangian atmospheric model with orography. Mon. Wea. Rev., 126 , 747–771.
Semazzi, F. H. M., J. H. Qian, and J. S. Scroggs, 1995: A global nonhydrostatic semi-Lagrangian atmospheric model without orography. Mon. Wea. Rev., 123 , 2534–2550.
Skamarock, W. C., and J. B. Klemp, 1992: The stability of time-split numerical methods for the hydrostatic and the nonhydrostatic elastic equations. Mon. Wea. Rev., 120 , 2109–2127.
Skamarock, W. C., and J. B. Klemp, . 1994: Efficiency and accuracy of the Klemp–Wilhemson time-splitting technique. Mon. Wea. Rev., 122 , 2623–2630.
Smolarkiewicz, P. K., V. Grubisic, L. G. Margolin, and A. A. Wyszogrodzki, 1999: Forward-in-time differencing for fluids: Non-hydrostatic modeling of fluid motions on a sphere. Proc. Seminar on Recent Developments in Numerical Methods for Atmospheric Modelling, Reading, United Kingdom, ECMWF, 21–43.
Straka, J. M., R. B. Wilhemson, L. J. Wicker, J. R. Anderson, and K. K. Droegemeier, 1993: Numerical solutions of a nonlinear density current: A benchmark solution and comparisons. Int. J. Numer. Methods Fluids, 17 , 1–22.
Taylor, K. E., 1984: A vertical finite-difference scheme for hydrostatic and nonhydrostatic equations. Mon. Wea. Rev., 112 , 1398–1402.
Tompkins, A. M., and G. C. Craig, 1998a: Time-scales of adjustment to radiative–convective equilibrium in the troposphere. Quart. J. Roy. Meteor. Soc., 124 , 2693–2713.
Tompkins, A. M., and G. C. Craig, . 1998b: Radiative-convective equilibrium in a three-dimensional cloud ensemble model. Quart. J. Roy. Meteor. Soc., 124 , 2073–2097.
Tompkins, A. M., and G. C. Craig, . 1999: Sensitivity of tropical convection to sea surface temperature in the absence of large-scale flow. J. Climate, 12 , 462–476.
Xue, M., K. K. Droegemeier, and V. Wong, 2000: The Advanced Regional Prediction System (ARP-S)—A multi-scale nonhydrostatic atmospheric simulation and prediction model. Part I: Model dynamics and verification. Meteor. Atmos. Phys., 75 , 161–193.
APPENDIX
Stability Analysis of the Implicit Scheme
In the scheme presented in section 2, the buoyancy term is also counted as an implicit term. This is for the consistency of the deformation of sound waves due to the stratification. In this appendix, we show the stability of the implicit scheme for sound and gravity waves in the isothermal atmosphere.





















Vertical profiles of perturbations of pressure p′ for the experiments of the vertical propagation of sound waves at (a) t = 10, (b) 20, and (c) 30 s. The initial state is shown by the dotted curve and the experiments with Δt = 0.1, 1.0, and 10.0 s are shown by the solid, dashed, and dashed–dotted curves, respectively
Citation: Monthly Weather Review 130, 5; 10.1175/1520-0493(2002)130<1227:CSFTCN>2.0.CO;2

Vertical profiles of perturbations of pressure p′ for the experiments of the vertical propagation of sound waves at (a) t = 10, (b) 20, and (c) 30 s. The initial state is shown by the dotted curve and the experiments with Δt = 0.1, 1.0, and 10.0 s are shown by the solid, dashed, and dashed–dotted curves, respectively
Citation: Monthly Weather Review 130, 5; 10.1175/1520-0493(2002)130<1227:CSFTCN>2.0.CO;2
Vertical profiles of perturbations of pressure p′ for the experiments of the vertical propagation of sound waves at (a) t = 10, (b) 20, and (c) 30 s. The initial state is shown by the dotted curve and the experiments with Δt = 0.1, 1.0, and 10.0 s are shown by the solid, dashed, and dashed–dotted curves, respectively
Citation: Monthly Weather Review 130, 5; 10.1175/1520-0493(2002)130<1227:CSFTCN>2.0.CO;2

The same as in Fig. 1 but for the vertical velocity w
Citation: Monthly Weather Review 130, 5; 10.1175/1520-0493(2002)130<1227:CSFTCN>2.0.CO;2

The same as in Fig. 1 but for the vertical velocity w
Citation: Monthly Weather Review 130, 5; 10.1175/1520-0493(2002)130<1227:CSFTCN>2.0.CO;2
The same as in Fig. 1 but for the vertical velocity w
Citation: Monthly Weather Review 130, 5; 10.1175/1520-0493(2002)130<1227:CSFTCN>2.0.CO;2

Time sequences of energies for the experiments of the vertical propagation of sound waves with Δt = 1 s. (a) Internal energy ρe (solid), potential energy ρΦ (dashed), and available potential energy ρ(e + Φ) (dotted). (b), (c), (d) Kinetic energy ρw2/2 (solid), available potential energy ρ(e + Φ) (dashed), and total energy ρetot (dotted). (a), (b) The noncorrection method, (c) the correction method, and (d) the conservative method (see text). Values of the energies are the differences from the initial values and averaged for a unit volume: the dimension is J m−3
Citation: Monthly Weather Review 130, 5; 10.1175/1520-0493(2002)130<1227:CSFTCN>2.0.CO;2

Time sequences of energies for the experiments of the vertical propagation of sound waves with Δt = 1 s. (a) Internal energy ρe (solid), potential energy ρΦ (dashed), and available potential energy ρ(e + Φ) (dotted). (b), (c), (d) Kinetic energy ρw2/2 (solid), available potential energy ρ(e + Φ) (dashed), and total energy ρetot (dotted). (a), (b) The noncorrection method, (c) the correction method, and (d) the conservative method (see text). Values of the energies are the differences from the initial values and averaged for a unit volume: the dimension is J m−3
Citation: Monthly Weather Review 130, 5; 10.1175/1520-0493(2002)130<1227:CSFTCN>2.0.CO;2
Time sequences of energies for the experiments of the vertical propagation of sound waves with Δt = 1 s. (a) Internal energy ρe (solid), potential energy ρΦ (dashed), and available potential energy ρ(e + Φ) (dotted). (b), (c), (d) Kinetic energy ρw2/2 (solid), available potential energy ρ(e + Φ) (dashed), and total energy ρetot (dotted). (a), (b) The noncorrection method, (c) the correction method, and (d) the conservative method (see text). Values of the energies are the differences from the initial values and averaged for a unit volume: the dimension is J m−3
Citation: Monthly Weather Review 130, 5; 10.1175/1520-0493(2002)130<1227:CSFTCN>2.0.CO;2

Distributions of perturbations of potential temperature for the experiment of the horizontal propagation of gravity waves at (a) the initial state and (b) t = 3000 s. The contour intervals are (a) 10−3 K and (b) 0.5 × 10−3 K. The solid curves are positive, the dashed curves are negative, and the dotted curves are zero
Citation: Monthly Weather Review 130, 5; 10.1175/1520-0493(2002)130<1227:CSFTCN>2.0.CO;2

Distributions of perturbations of potential temperature for the experiment of the horizontal propagation of gravity waves at (a) the initial state and (b) t = 3000 s. The contour intervals are (a) 10−3 K and (b) 0.5 × 10−3 K. The solid curves are positive, the dashed curves are negative, and the dotted curves are zero
Citation: Monthly Weather Review 130, 5; 10.1175/1520-0493(2002)130<1227:CSFTCN>2.0.CO;2
Distributions of perturbations of potential temperature for the experiment of the horizontal propagation of gravity waves at (a) the initial state and (b) t = 3000 s. The contour intervals are (a) 10−3 K and (b) 0.5 × 10−3 K. The solid curves are positive, the dashed curves are negative, and the dotted curves are zero
Citation: Monthly Weather Review 130, 5; 10.1175/1520-0493(2002)130<1227:CSFTCN>2.0.CO;2

The same as Figs. 3a and 3b but for the experiment of the horizontal propagation of gravity waves. In (b), the total kinetic energy ρ(u2 + w2)/2 (solid), ρu2/2 and ρw2/2 (dashed–dotted) are added
Citation: Monthly Weather Review 130, 5; 10.1175/1520-0493(2002)130<1227:CSFTCN>2.0.CO;2

The same as Figs. 3a and 3b but for the experiment of the horizontal propagation of gravity waves. In (b), the total kinetic energy ρ(u2 + w2)/2 (solid), ρu2/2 and ρw2/2 (dashed–dotted) are added
Citation: Monthly Weather Review 130, 5; 10.1175/1520-0493(2002)130<1227:CSFTCN>2.0.CO;2
The same as Figs. 3a and 3b but for the experiment of the horizontal propagation of gravity waves. In (b), the total kinetic energy ρ(u2 + w2)/2 (solid), ρu2/2 and ρw2/2 (dashed–dotted) are added
Citation: Monthly Weather Review 130, 5; 10.1175/1520-0493(2002)130<1227:CSFTCN>2.0.CO;2

Distributions of perturbations of (a) horizontal velocity and (b) vertical velocity for the experiment of the finite amplitude topographic waves at t = 3000 s. The contour intervals are (a) 6 and (b) 0.6 m s−1. The solid curves are positive, the dashed curves are negative, and the dotted curves are zero
Citation: Monthly Weather Review 130, 5; 10.1175/1520-0493(2002)130<1227:CSFTCN>2.0.CO;2

Distributions of perturbations of (a) horizontal velocity and (b) vertical velocity for the experiment of the finite amplitude topographic waves at t = 3000 s. The contour intervals are (a) 6 and (b) 0.6 m s−1. The solid curves are positive, the dashed curves are negative, and the dotted curves are zero
Citation: Monthly Weather Review 130, 5; 10.1175/1520-0493(2002)130<1227:CSFTCN>2.0.CO;2
Distributions of perturbations of (a) horizontal velocity and (b) vertical velocity for the experiment of the finite amplitude topographic waves at t = 3000 s. The contour intervals are (a) 6 and (b) 0.6 m s−1. The solid curves are positive, the dashed curves are negative, and the dotted curves are zero
Citation: Monthly Weather Review 130, 5; 10.1175/1520-0493(2002)130<1227:CSFTCN>2.0.CO;2

The same as Figs. 3a and 3b but for the experiment of the finite amplitude topographic waves. In (b), energy change due to the damping terms ER (dashed–dotted) and the difference between total energy ρetot and ER (dotted) are added
Citation: Monthly Weather Review 130, 5; 10.1175/1520-0493(2002)130<1227:CSFTCN>2.0.CO;2

The same as Figs. 3a and 3b but for the experiment of the finite amplitude topographic waves. In (b), energy change due to the damping terms ER (dashed–dotted) and the difference between total energy ρetot and ER (dotted) are added
Citation: Monthly Weather Review 130, 5; 10.1175/1520-0493(2002)130<1227:CSFTCN>2.0.CO;2
The same as Figs. 3a and 3b but for the experiment of the finite amplitude topographic waves. In (b), energy change due to the damping terms ER (dashed–dotted) and the difference between total energy ρetot and ER (dotted) are added
Citation: Monthly Weather Review 130, 5; 10.1175/1520-0493(2002)130<1227:CSFTCN>2.0.CO;2

Distributions of perturbations of potential temperature for the cold-bubble experiment with the grid interval Δx = Δz = 50 m at (a) the initial state, (b) t = 300, (c) 600, and (d) 900 s. The contour intervals are 1 K
Citation: Monthly Weather Review 130, 5; 10.1175/1520-0493(2002)130<1227:CSFTCN>2.0.CO;2

Distributions of perturbations of potential temperature for the cold-bubble experiment with the grid interval Δx = Δz = 50 m at (a) the initial state, (b) t = 300, (c) 600, and (d) 900 s. The contour intervals are 1 K
Citation: Monthly Weather Review 130, 5; 10.1175/1520-0493(2002)130<1227:CSFTCN>2.0.CO;2
Distributions of perturbations of potential temperature for the cold-bubble experiment with the grid interval Δx = Δz = 50 m at (a) the initial state, (b) t = 300, (c) 600, and (d) 900 s. The contour intervals are 1 K
Citation: Monthly Weather Review 130, 5; 10.1175/1520-0493(2002)130<1227:CSFTCN>2.0.CO;2

The same as Fig. 8d but for the grid interval with (a) 100 and (b) 200 m
Citation: Monthly Weather Review 130, 5; 10.1175/1520-0493(2002)130<1227:CSFTCN>2.0.CO;2

The same as Fig. 8d but for the grid interval with (a) 100 and (b) 200 m
Citation: Monthly Weather Review 130, 5; 10.1175/1520-0493(2002)130<1227:CSFTCN>2.0.CO;2
The same as Fig. 8d but for the grid interval with (a) 100 and (b) 200 m
Citation: Monthly Weather Review 130, 5; 10.1175/1520-0493(2002)130<1227:CSFTCN>2.0.CO;2

The same as Figs. 3a and 3b but for the cold-bubble experiment. (a), (b) The correction method, and (c) the conservative method (see text)
Citation: Monthly Weather Review 130, 5; 10.1175/1520-0493(2002)130<1227:CSFTCN>2.0.CO;2

The same as Figs. 3a and 3b but for the cold-bubble experiment. (a), (b) The correction method, and (c) the conservative method (see text)
Citation: Monthly Weather Review 130, 5; 10.1175/1520-0493(2002)130<1227:CSFTCN>2.0.CO;2
The same as Figs. 3a and 3b but for the cold-bubble experiment. (a), (b) The correction method, and (c) the conservative method (see text)
Citation: Monthly Weather Review 130, 5; 10.1175/1520-0493(2002)130<1227:CSFTCN>2.0.CO;2

Fig. A1. Dependencies of the growth rate |λ| on the Courant number ν and the implicit factor α
Citation: Monthly Weather Review 130, 5; 10.1175/1520-0493(2002)130<1227:CSFTCN>2.0.CO;2

Fig. A1. Dependencies of the growth rate |λ| on the Courant number ν and the implicit factor α
Citation: Monthly Weather Review 130, 5; 10.1175/1520-0493(2002)130<1227:CSFTCN>2.0.CO;2
Fig. A1. Dependencies of the growth rate |λ| on the Courant number ν and the implicit factor α
Citation: Monthly Weather Review 130, 5; 10.1175/1520-0493(2002)130<1227:CSFTCN>2.0.CO;2