1. Introduction
The composition of the atmosphere below ∼100-km altitude remains roughly constant due to turbulent mixing. This affords useful simplifications to the atmospheric fluid equations that can improve computational speed and efficiency of the discretized forms used as dynamical cores for operational numerical weather prediction (NWP).
As NWP models are progressively extended upward, these convenient lower atmospheric simplifications no longer apply (Akmaev 2011). Kinematic molecular viscosities increase approximately exponentially with height such that, above ∼100-km altitude, they become large enough to suppress turbulence. Absent turbulent mixing, thermospheric composition varies with height according to species mass. Photolysis also leads to atoms ultimately dominating molecules as the major mass component of the thermosphere. These changes cause the specific gas constant R, and the mass specific heats at constant pressure and volume, cp and cυ, respectively, to increase with height in the thermosphere and to vary in space and time in response to changes in species concentrations. Likewise, the monotonic decrease in gravitational acceleration g with height yields significantly smaller thermospheric values relative to the surface.
Modifications to a model’s dynamical core to allow these parameters to vary will typically apply at all altitudes. If extensive intrusive changes are involved, such modifications may present barriers to operational transition. For example, additional numerical complexity and overhead can jeopardize continued reliable delivery of tropospheric forecasts within inflexible time windows using the finite computational resources available at operational centers. Indeed, to forecast R, cp, and cυ as global time-varying state variables ideally requires new initialization, photochemistry, and transport code for a range of new thermospheric constituents. In addition to adding considerable new computational overhead, there are at present no operational observations of thermospheric constituents to assimilate as initial conditions and verification for these constituent forecasts, suggesting both questionable and largely unverifiable thermospheric skill impacts.
With these realities in mind, while also recognizing new needs to extend NWP models into the thermosphere to support space weather applications (e.g., Jackson et al. 2019; Berger et al. 2020), here we seek simpler and less intrusive dynamical core upgrades that have minimal impacts on how existing NWP dynamical cores are formulated and run in the lower atmosphere, yet capture salient aspects of variable composition thermospheric dynamics. While our development here focuses on the discretized hydrostatic primitive equations (HPEs) within the Navy Global Environmental Model (NAVGEM), as discussed in section 2, our hybrid virtual potential temperature concepts, developed in section 3, also apply to the many other atmospheric models using virtual potential temperature formulations, including nonhydrostatic models. Variable gravitational acceleration is briefly considered in section 4. The new dynamical core is tested and numerically verified in section 5 for two new idealized states representing hot thermospheres, one incorporating variable R and κ = R/cp. Results are summarized in section 6.
2. NAVGEM governing equations
a. Equations of state and hydrostatic balance
b. Pressure gradient force
3. Variable composition and new virtual potential temperature forms
a. Variable gas constants
Solid curves in Fig. 1a show representative profiles of R, κ−1, and cp as a function of the solar 10.7 cm solar radio flux (F10.7) index for atmospheric pressures extending from the surface (1000 hPa) to near the exobase at ∼500 km (10−8 hPa), based on calculations and parameterized fits described in the appendix. We see in Fig. 1a that R increases by up to 100% in the thermosphere relative to its constant lower atmospheric value. Corresponding changes in cp in Fig. 1c are lower, due to the accompanying decrease in κ−1 with height in Fig. 1b.
While (18)–(20b) are bulk terms involving all mass constituents, in what follows we will split and separate the tropospheric and thermospheric contributions, since they are widely separated in altitude and thus independent of one another. This separation also helps to illustrate how the new thermospheric contributions to virtual temperature relate to the original moist tropospheric terms.
b. Moist virtual potential temperature (MVPT)
The key point here is that by virtualizing our prognostic potential temperature variable
While for simplicity we have considered only the gas phase here, it is straightforward to include additional moist modifications due to liquid water and ice (e.g., Davies et al. 2005).
c. Thermospheric virtual potential temperature
While we only show vertical profiles of thermospheric Rt in Fig. 1a, we note that the virtual thermospheric temperature (28) and the hybrid forms that follow all hold for an Rt that can vary in all three spatial dimensions and in time, just like sw in (24).
d. Hybrid virtual potential temperature (HVPT)
e. Mass specific heats using HVPT
f. Variable κ: A corrected HVPT
Next, we explore building variable κ into a θ-based dynamical core using a corrected HVPT variable.
1) Specific enthalpy
2) Corrected HVPT
The major change to the virtual potential temperature definition in (47) is the
A new Dκ/Dt term appears in (49), needed to remove an identical term contained within the material derivative of
The appearance of Dκ/Dt terms in (49) is unavoidable since potential temperature is no longer conserved in variable-κ atmospheres. Evaluating Dκ/Dt generally requires solutions from additional prognostic equations governing constituent transport and photochemistry (e.g., Juang 2011; Liu et al. 2018), which are nontrivial and potentially expensive additions. For our one-dimensional profile on isobaric η surfaces in Fig. 1b, Dκ/Dt simplifies to ω(∂κ/∂p), where ∂κ/∂p is given by the analytical p derivative of the fitted expression (A3) (if F10.7 varies during the forecast, a ∂κ/∂t term is also required).
3) Corrected potential temperature and potential enthalpy
It proves useful at this point to consider briefly a closely related class of NWP models that use potential temperature θ = TΠ−1 as the prognostic variable, then derive
4) Standard HVPT errors due to constant
What other typical errors do we anticipate when using a globally constant
The specific κ dependences in (55) and (57) reduce the impacts of thermospheric κ biases on these key wave parameters. For example, reading off κ values from Fig. 1b, the high bias in Cs from using constant
The effect of κ bias on N is diminished by noting that the term in square parentheses in (55) is dominated by ∂T/∂Z, which is large and positive in the thermosphere, with the κ-dependent lapse rate much smaller in magnitude by comparison. Errors become larger in the upper thermosphere, where ∂T/∂Z → 0 and κ departures from
Compare these errors to those from omitting temperature virtualization and assuming constant R, which scale as R1/2 for both N and Cs. Reading off R values from Fig. 1a, they are ∼21% at 10−6 hPa and ∼36% at 10−8 hPa, about 4–6 times higher than the corresponding relative errors in Cs from assuming constant κ. These errors are in addition to the huge errors in geopotential and pressure gradient force that result from using constant R in the thermosphere, as previously discussed.
Thus, we conclude that an upgraded HVPT-based dynamical core represents an adequate bridging strategy to allow an existing dynamical core to be extended into the thermosphere to capture realistic composition-induced changes to the equations of state, hydrostatic balance, and momentum balance, with acceptable thermodynamic error biases due to use of constant κ. In cases where accurate κ distributions are required in the forecast, such as deeper vertical extensions into the thermosphere, the corrected HVPT form (49) or its locally constant approximation (52) can be employed.
g. Diabatic tendency Q
We now briefly consider the right-hand sides of the HVPT Eqs. (39) and (49), where diabatic heating due to physics is represented as a net energy tendency Q. Since physics terms are more naturally applied in NAVGEM as temperature tendencies, then after the dynamics step and immediately before the physics update, we convert HVPT (either
The NWP community has revisited thermodynamic potentials as a more fundamental basis for self-consistently deriving thermodynamic equations that capture latent heat terms due to phase changes in water within the dynamical core (Thuburn 2017; Staniforth and White 2019). In this work we have assumed for simplicity that latent heating rates are encapsulated within Q. Composition changes due to ion-neutral photochemistry can also change specific thermodynamic potentials such as enthalpy [see section 2.2 of Eldred et al. (2022)]. Again, for simplicity we have made no attempt here to extract any additional composition-dependent chemical potential terms for explicit inclusion in the specific enthalpy or HVPT equations, assuming instead that any heating due to thermospheric composition changes is also encapsulated within Q. For example, heating of the mesosphere and thermosphere due to exothermic photochemical reactions that alter composition are typically parameterized as a direct diabatic contribution to Q (e.g., Marsh et al. 2007). Despite these differences, our investigation of errors due to a constant-κ prognostic virtual potential temperature variable has initiated a similarly motivated preliminary error assessment for the thermosphere to those deeper tropospheric assessments based on thermodynamic potentials [cf. sections 3 and 4 of Eldred et al. (2022)].
4. Variable gravitational acceleration
We now briefly consider constant versus variable g. Use of constant g is generally understood to follow from the shallow-atmosphere approximation used to derive the HPEs (White et al. 2005). In the lower atmosphere the errors from such an approximation are self-evidently small. For example, scale analyses suggest other terms removed by the shallow-atmosphere approximation, such as cosφ Coriolis terms, yield greater errors by comparison (White and Bromley 1995).
In his influential review paper on whole-atmosphere modeling, Akmaev (2011) challenged the need for constant-g within ground-to-thermosphere models using the HPEs. His arguments focus on the unusual property of the HPEs, stemming from their use of pressure as the vertical coordinate, that g disappears from the discretized equations, being subsumed via hydrostatics into a temperature-dependent Φ given in (12). Akmaev (2011) argued that the absence of g in the HPEs implies that their prognostic dynamics are (quote) “formally independent of whether or not g is constant.” Paraphrasing his arguments, the HPEs provide solutions on model pressure surfaces that are valid for any choice of g(z), subject only to the ΦS lower boundary condition in (12). Relatedly, Akmaev (2011) also rejected arguments that variable g violates the shallow-atmosphere approximation (see, e.g., section 3b of White et al. 2005), arguing that the choice for g depends only on hydrostatics that exist largely separate from and independent of the shallow-atmosphere approximations.
One can immediately see from (62) that one can only proceed to the pure pressure-tendency form of the continuity equation as used in the HPEs by setting g0/g = 1, consistent with the constant-g approximation routinely adopted within HPE-based models.
This provides a simple proof of the assertion of Deng et al. (2008) that introducing variable g(z) into HPE-based models introduces an inconsistency in mass-wind balance, recognizing that this continuity equation is used subsequently to infer pressure tendencies and associated vertical motion. Put another way, (62) disproves the assertion of Akmaev (2011) that the output of models derived from the HPEs is formally independent of any assumed form for g. In terms of the shallow-atmosphere approximation, White and Wood (2012), extending the earlier work of White et al. (2005), argued that constant g is required in the shallow-atmosphere approximation in order to avoid a spurious nonzero net divergence of the gravitational acceleration vector. Indeed, White and Wood (2012) show that meridional variations in g due to departures from a spherical Earth are in some ways easier to accommodate within the HPEs than vertical variations.
Our long-term strategy is to transition from NAVGEM to a new forecast model incorporating nonhydrostatic deep-atmosphere forms of the discretized equations that allow variable g to be implemented directly into the core (e.g., Wood et al. 2014).
5. Tests in NAVGEM using idealized globally balanced states
We augmented the dynamical core in NAVGEM with options to run using either MVPT, standard HVPT, or corrected HVPT, then extended the vertical range to ∼200-km altitude (ptop = 5 × 10−7 hPa) in a 160 level (L160) formulation and to ∼400-km altitude (ptop = 5 × 10−10 hPa) as an L184 configuration.
A range of idealized dynamical core experiments was performed to test the accuracy and robustness of these new vertically extended configurations using the new HVPT-based dynamical core. We present two examples that prescribe an idealized globally balanced state as atmospheric initial conditions. The model is tested for its ability to maintain that balanced state over a 5-day forecast integration.
a. Variable-R thermosphere-like balanced states on nonrotating planet
1) Analytical solutions
This balanced solution provides an excellent test of the ability of HVPT dynamics to maintain a globally balanced thermospheric wind field in which both thermospheric temperature T(z) and specific gas constant R(z) can vary in the vertical, subject to the constraints that (i) the integral in (72a) must exist, and (ii) p → 0 as z → ∞. Although κ does not appear in these solutions, it can be shown that the compatibility relations (66) and (67) and their ensuing balanced solutions hold for atmospheres with any sufficiently smooth vertical variation in κ.
2) Implementation and tests in NAVGEM
Because these solutions are formulated in geometric height rather than pressure, we implement this solution as a NAVGEM initial condition as follows. We use (72b) to initialize pS, then use gridpoint pressures p(η) on the left hand side of (72a) to solve (72a) iteratively for a corresponding initial z(η), and hence an initial geopotential Φ = g0z, using Newton’s method with known initial profiles R(z) and T(z). These z(η) values in turn initialize the model temperature T(z), specific gas constants R(z), and zonal winds U(φ, z) from (72c).
To suppress initialization noise through errors in these iterative numerical procedures, we choose analytical functional forms for R(z) and T(z) that provide an analytical solution to the height integral in (72a). We choose piecewise fits for R(z) and T(z), shown in Fig. 3, such that 1/[R(z)T(z)] is quadratic in z above 100 km, and thus analytically integrable. Note in Fig. 3 that the virtual thermospheric temperature
To test the ability of the standard HVPT core in NAVGEM to maintain this variable-R balanced state, we set
The same experiment was repeated using the deeper T119L184 configuration using different vertical profiles for R and T to prevent values from becoming excessively large. The results (not shown) also maintained the analytical balanced wind and temperature states out to 120 h, with largest errors confined to the uppermost few model layers.
As a corresponding test of the ability of the corrected HVPT core in NAVGEM to maintain this state, we allowed κ to vary vertically as
b. Globally balanced state on rotating planet with thermospheric heating
This test is designed to establish a balanced tropospheric state, then to apply intense thermosphere-like heating above it as the forecast proceeds, to assess the ability of the core to maintain both states at the lower and upper boundaries. As opposed to the previous case, planetary rotation is included.
Thus, this idealized experiment establishes a balanced Ullrich et al. (2014) state in the lower atmosphere from the surface to 10−2 hPa, while applying rapid thermosphere-like heating from 10−4 to 10−6 hPa during the forecast. The goal is to see whether the core maintains the balanced tropospheric state while simultaneously maintaining a motionless balanced thermosphere undergoing rapid heating of the isothermal profile. This heating experiment was implemented within NAVGEM and integrated out to five days (120 h) at T119L160 using α = 0.168 56 h−1 (∼6-h e-folding time). The forecast thermosphere is heated to within 1% of the limiting 700-K isothermal state after ∼27.3 h.
The 0-h balanced initial wind fields are plotted in the top row of Fig. 7. Despite imposition of intense thermospheric heating as soon as the forecast commences, these initially balanced winds are maintained throughout the forecast up to high altitudes, where zonal-mean deviations from these 0-h initial conditions after 120 h are <1 m s−1 at all heights and latitudes, and ≪1 m s−1 at heights below the uppermost few model layers.
These experiments verify that the model maintains a balanced circulation containing a representative idealized troposphere capped by an intensely heated thermosphere. It should be noted that the elastic upper boundary condition of our pressure-based dynamical core allows the thermosphere to expand naturally (via diagnostic increases in
6. Summary and outlook
We have presented a methodology, based around the classical NWP concept of virtual temperature, that allows us to take a preexisting dynamical core designed solely for lower-atmospheric NWP and, via targeted changes that are minimally invasive lower down, to extend it through the thermosphere to capture salient aspects of the variable composition dynamics of the thermosphere. The augmentation is structured around a hybrid virtual potential temperature (HVPT) that asymptotically approaches moist and thermospheric limits in the lower and upper atmospheres, respectively. We implemented these HVPT-based upgrades within the NAVGEM dynamical core, then tested and validated both the standard and corrected HVPT dynamical-core formulations using forecasts of test cases incorporating idealized balanced initial conditions, hot or rapidly heated thermospheres, and height-varying R and κ in the thermosphere.
We are not aware of any previous attempts to generalize virtual temperature to allow lower atmospheric dynamical cores to extend through the thermosphere. One reason could be that moist virtual temperature is a small modification, typically within a percent of dry air temperatures even within highly moist tropospheric environments. By contrast, thermospheric virtual temperatures can be ∼100% larger than local air temperatures in the upper thermosphere. Our work shows that there is no small-perturbation restriction to the virtualization approach, permitting a generalized HVPT that incorporates large changes in thermospheric composition, relative to the lower atmosphere, in ways that accurately capture impacts on adiabatic dynamics throughout the thermosphere.
We consider a major practical advantage of the proposed HVPT modification to be its light footprint on existing tropospheric numerics. The HVPT modification returns the exact MVPT-based equations used for tropospheric NWP by ensuring that
Consistent with our goal of fast-tracking thermospheric NWP development, these and other promising initial findings led us to implement the HVPT dynamical core in a full physics version of NAVGEM running at both T119L160 and T119L184, to provide a background forecast model to accelerate development of NWP-quality thermospheric physics as well as ground-to-thermosphere ensemble data assimilation capabilities within a mature cycling NWP system. This was motivated by an emerging need to drive physics-based ionospheric models with accurate thermospheric forecasts as inputs (e.g., Zawdie et al. 2020). This work will be described more fully elsewhere, but we can report that our prototype HVPT-based forecast model has remained stable and accurate to date, operating now for well over a year at both T119L160 and T119L184 in a variety of research experiments. These include cycling ensemble-based data assimilation experiments extending over months, as well as free-running full-physics nature runs out to 40 days and longer. Thus, our HVPT-based development described here is already achieving its primary goal of fast-tracking ground-to-thermosphere NWP development, as we await a next-generation nonhydrostatic spectral-element forecast model, slated to replace NAVGEM for NWP and currently under active research and development (e.g., Theurich et al. 2016; Zaron et al. 2022). We are concurrently developing high-altitude capabilities for that new model based on new deep-atmosphere nonhydrostatic equation sets without any inbuilt assumptions about composition to admit variable R, κ, and g directly within the dynamical core.
Our validated virtual temperature concepts for the thermosphere also offer promising scientific research diagnostics for understanding thermospheric dynamics, particularly when analyzing zonal-mean or planetary-scale circulations for which constant-κ errors due to vertical motion are small. We leave this for future work, but note a few possibilities, such as thermal-wind and gradient-wind relations based on thermospheric virtual temperature (28) (see, e.g., White and Staniforth 2008), and a thermospheric potential vorticity based on thermospheric virtual potential temperature (see, e.g., Schubert et al. 2001; Schubert 2004; White et al. 2005).
Acknowledgments.
This work was supported by the Defense Sciences Office of the Defense Advanced Research and Projects Agency (DARPA DSO) through their Space Environment Exploitation (SEE) program.
Data availability statement.
IDL code containing all data needed to create the fitted R, cp,and κ profiles as described in the appendix and to reproduce and plot the profiles in Fig. 1 is available at https://map.nrl.navy.mil/map/pub/nrl/papers/hvpt/. NAVGEM idealized model output data are available from the authors upon request. MSIS data and code are available at https://map.nrl.navy.mil/map/pub/nrl/NRLMSIS/.
APPENDIX
Representative R and κ Profiles for Use in NAVGEM
Since the standard MVPT-based dynamical core adopts constant R and cp, we sought simple robust initial parameterizations of vertical variations of R and cp that captured the basic geophysical dependences needed for more realistic forecasts of ground-to-thermosphere dynamics using a vertically extended NAVGEM with a HVPT-based dynamical core. These fits comprise a simple intermediate step prior to implementing more detailed code upgrades to incorporate prognostic thermospheric composition, as required to compute R and cp values directly.
a. Specific gas constant R
We computed R profiles using temperature and composition from the Mass Spectrometer Incoherent Scatter (MSIS) empirical model (Picone et al. 2002) in a global-mean configuration with diurnal variations included. Examples for a solar F10.7 = 100 solar flux units (sfu) and geomagnetic Ap = 4 on day 80 (April) at the equator and Greenwich meridian are shown in Fig. A1a, comprising six profiles at equi-spaced universal times spanning a diurnal cycle. As opposed to corresponding profiles plotted versus geometric height (not shown), these pressure-level profiles do not vary appreciably with local time, indicating that diurnal variations result mostly from atmospheric thermal expansion on the hot dayside and contraction on the cold nightside.
b. Specific heats cp and cυ via κ−1 fits
We utilize the separable form (20) to fit κ−1, which then gives cp and cυ by scaling the fitted R profile above. In choosing an appropriate analytical function to fit κ−1, we note that the deep atmosphere is characterized by the diatomic limit
c. Solar cycle variability
These profiles and their associated fitting coefficients are sensitive to solar activity. Solid curves in Fig. A2 show how these coefficients vary with F10.7. Thicker transparent curves show third-order polynomial fits, which capture the salient F10.7 variations very well for the
Coefficients pj corresponding to cubic polynomial fits of the form
REFERENCES
Akmaev, R. A., 2011: Whole atmosphere modeling: Connecting terrestrial and space weather. Rev. Geophys., 49, RG4004, https://doi.org/10.1029/2011RG000364.
Akmaev, R. A., and H.-M. H. Juang, 2008: Using enthalpy as a prognostic variable in atmospheric modelling with variable composition. Quart. J. Roy. Meteor. Soc., 134, 2193–2197, https://doi.org/10.1002/qj.345.
Arakawa, A., and M. J. Suarez, 1983: Vertical differencing of the primitive equations in sigma-coordinates. Mon. Wea. Rev., 111, 34–45, https://doi.org/10.1175/1520-0493(1983)111%3C0034:VDOTPE%3E2.0.CO;2.
Berger, T. E., M. J. Holzinger, E. K. Sutton, and J. P. Thayer, 2020: Flying through uncertainty. Space Wea., 18, e2019SW002373, https://doi.org/10.1029/2019SW002373.
Davies, T., A. Staniforth, N. Wood, and J. Thuburn, 2003: Validity of anelastic and other equation sets as inferred from normal-mode analysis. Quart. J. Roy. Meteor. Soc., 129, 2761–2775, https://doi.org/10.1256/qj.02.1951.
Davies, T., M. J. P. Cullen, A. J. Malcolm, M. H. Mawson, A. Staniforth, A. A. White, and N. Wood, 2005: A new dynamical core for the Met Office’s global and regional modelling of the atmosphere. Quart. J. Roy. Meteor. Soc., 131, 1759–1782, https://doi.org/10.1256/qj.04.101.
Deng, Y., A. J. Ridley, and W. Wang, 2008: Effect of the altitudinal variation of the gravitational acceleration on the thermosphere simulation. J. Geophys. Res., 113, A09302, https://doi.org/10.1029/2008JA013081.
Eckermann, S. D., 2009: Hybrid σ–p coordinate choices for a global model. Mon. Wea. Rev., 137, 224–245, https://doi.org/10.1175/2008MWR2537.1.
Eckermann, S. D., and Coauthors, 2009: High-altitude data assimilation system experiments for the northern summer mesosphere season of 2007. J. Atmos. Sol.-Terr. Phys., 71, 531–551, https://doi.org/10.1016/j.jastp.2008.09.036.
Eckermann, S. D., and Coauthors, 2018: High-altitude (0–100 km) global atmospheric reanalysis system: Description and application to the 2014 austral winter of the Deep Propagating Gravity Wave Experiment (DEEPWAVE). Mon. Wea. Rev., 146, 2639–2666, https://doi.org/10.1175/MWR-D-17-0386.1.
Eldred, C., M. Taylor, and O. Guba, 2022: Thermodynamically consistent versions of approximations used in modelling moist air. Quart. J. Roy. Meteor. Soc., 148, 3184–3210, https://doi.org/10.1002/qj.4353.
Haltiner, G. J., and R. T. Williams, 1980: Numerical Weather Prediction and Dynamic Meteorology. 2nd ed. Wiley, 477 pp.
Hodur, R. M., 1997: The Naval Research Laboratory’s Coupled Ocean/Atmosphere Mesoscale Prediction System (COAMPS). Mon. Wea. Rev., 125, 1414–1430, https://doi.org/10.1175/1520-0493(1997)125%3C1414:TNRLSC%3E2.0.CO;2.
Hogan, T. F., and T. E. Rosmond, 1991: The description of the Navy operational global atmospheric prediction system’s spectral forecast model. Mon. Wea. Rev., 119, 1786–1815, https://doi.org/10.1175/1520-0493(1991)119%3C1786:TDOTNO%3E2.0.CO;2.
Hogan, T. F., and Coauthors, 2014: The Navy Global Environmental Model. Oceanography, 27, 116–125, https://doi.org/10.5670/oceanog.2014.73.
Jackson, D. R., T. J. Fuller-Rowell, D. J. Griffin, M. J. Griffith, C. W. Kelly, D. R. Marsh, and M.-T. Walach, 2019: Future directions for whole atmosphere modeling: Developments in the context of space weather. Space Wea., 17, 1342–1350, https://doi.org/10.1029/2019SW002267.
Juang, H.-M. H., 2011: A multiconserving discretization with enthalpy as a thermodynamic prognostic variable in generalized hybrid vertical coordinates for the NCEP global forecast system. Mon. Wea. Rev., 139, 1583–1607, https://doi.org/10.1175/2010MWR3295.1.
Kasahara, A., 1974: Various vertical coordinate systems used for numerical weather prediction. Mon. Wea. Rev., 102, 509–522, https://doi.org/10.1175/1520-0493(1974)102%3C0509:VVCSUF%3E2.0.CO;2.
Kelly, J. F., S. Reddy, F. X. Giraldo, P. A. Reinecke, J. T. Emmert, M. Jones Jr., and S. D. Eckermann, 2023: A physics-based open atmosphere boundary condition for height-coordinate atmospheric models. J. Comput. Phys., 482, 112044, https://doi.org/10.1016/j.jcp.2023.112044.
Klemp, J. B., and R. B. Wilhelmson, 1978: Simulation of 3-dimensional convective storm dynamics. J. Atmos. Sci., 35, 1070–1096, https://doi.org/10.1175/1520-0469(1978)035%3C1070:TSOTDC%3E2.0.CO;2.
Klemp, J. B., and W. C. Skamarock, 2021: Adapting the MPAS dynamical core for applications extending into the thermosphere. J. Adv. Model. Earth Syst., 13, e2021MS002499, https://doi.org/10.1029/2021MS002499.
Klemp, J. B., and W. C. Skamarock, 2022: A constant pressure upper boundary formulation for models employing height-based vertical coordinates. Mon. Wea. Rev., 150, 2175–2186, https://doi.org/10.1175/MWR-D-21-0328.1.
Liu, H.-L., and Coauthors, 2018: Development and validation of the whole atmosphere community climate model with thermosphere and ionosphere extension (WACCM-X 2.0). J. Adv. Model. Earth Syst., 10, 381–402, https://doi.org/10.1002/2017MS001232.
Marsh, D. R., R. R. Garcia, D. E. Kinnison, B. A. Boville, F. Sassi, S. C. Solomon, and K. Matthes, 2007: Modeling the whole atmosphere response to solar cycle changes in radiative and geomagnetic forcing. J. Geophys. Res., 112, D23306, https://doi.org/10.1029/2006JD008306.
Maute, A., 2017: Thermosphere-ionosphere-electrodynamics general circulation model for the ionospheric connection explorer: TIEGCM-ICON. Space Sci. Rev., 212, 523–551, https://doi.org/10.1007/s11214-017-0330-3.
Mayne, N. J., and Coauthors, 2014: The unified model, a fully-compressible, non-hydrostatic, deep atmosphere global circulation model, applied to hot Jupiters. Astron. Astrophys., 561, A1, https://doi.org/10.1051/0004-6361/201322174.
Mendonça, J. M., S. L. Grimm, L. Grosheintz, and K. Heng, 2016: THOR: A new and flexible global circulation model to explore planetary atmospheres. Astrophys. J., 829, 115, https://doi.org/10.3847/0004-637X/829/2/115.
Picone, J. M., A. E. Hedin, D. P. Drob, and A. C. Aikin, 2002: NRLMSISE-00 empirical model of the atmosphere: Statistical comparisons and scientific issues. J. Geophys. Res., 107, 1468, https://doi.org/10.1029/2002JA009430.
Polichtchouk, I., S. Malardel, and M. Diamantakis, 2020: Potential temperature as a prognostic variable in hydrostatic semi-implicit semi-Lagrangian IFS. ECMWF Tech. Memo. 869, 22 pp., https://doi.org/10.21957/ss3tvs4yq.
Ritchie, H., C. Temperton, A. Simmons, M. Hortal, T. Davies, D. Dent, and M. Hamrud, 1995: Implementation of the semi-Lagrangian method in a high-resolution version of the ECMWF forecast model. Mon. Wea. Rev., 123, 489–514, https://doi.org/10.1175/1520-0493(1995)123%3C0489:IOTSLM%3E2.0.CO;2.
Schubert, W. H., 2004: A generalization of Ertel’s potential vorticity to a cloudy precipitating atmosphere. Meteor. Z., 13, 465–471, https://doi.org/10.1127/0941-2948/2004/0013-0465.
Schubert, W. H., S. A. Hausman, M. Garcia, K. V. Ooyama, and H.-C. Kuo, 2001: Potential vorticity in a moist atmosphere. J. Atmos. Sci., 58, 3148–3157, https://doi.org/10.1175/1520-0469(2001)058%3C3148:PVIAMA%3E2.0.CO;2.
Simmons, A. J., and R. Strüfing, 1983: Numerical forecasts of stratospheric warming events using a model with a hybrid vertical coordinate. Quart. J. Roy. Meteor. Soc., 109, 81–111, https://doi.org/10.1002/qj.49710945905.
Simmons, A. J., and C. Temperton, 1997: Stability of a two-time-level semi-implicit integration scheme for gravity wave motion. Mon. Wea. Rev., 125, 600–615, https://doi.org/10.1175/1520-0493(1997)125%3C0600:SOATTL%3E2.0.CO;2.
Skamarock, W. C., J. B. Klemp, M. G. Duda, L. D. Fowler, S.-H. Park, and T. D. Ringler, 2012: A multiscale nonhydrostatic atmospheric model using centroidal Voronoi tesselations and C-grid staggering. Mon. Wea. Rev., 140, 3090–3105, https://doi.org/10.1175/MWR-D-11-00215.1.
Skamarock, W. C., H. Ong, and J. B. Klemp, 2021: A fully compressible nonhydrostatic deep-atmosphere equations solver for MPAS. Mon. Wea. Rev., 149, 571–583, https://doi.org/10.1175/MWR-D-20-0286.1.
Staniforth, A., and A. White, 2019: Forms of the thermodynamic energy equation for moist air. Quart. J. Roy. Meteor. Soc., 145, 386–393, https://doi.org/10.1002/qj.3421.
Theurich, G., and Coauthors, 2016: The earth system prediction suite: Toward a coordinated U.S. modeling capability. Bull. Amer. Meteor. Soc., 97, 1229–1247, https://doi.org/10.1175/BAMS-D-14-00164.1.
Thuburn, J., 2017: Use of the Gibbs thermodynamic potential to express the equation of state in atmospheric models. Quart. J. Roy. Meteor. Soc., 143, 1185–1196, https://doi.org/10.1002/qj.3020.
Tremblin, P., and Coauthors, 2017: Advection of potential temperature in the atmosphere of irradiated exoplanets: A robust mechanism to explain radius inflation. Astrophys. J., 841, 30, https://doi.org/10.3847/1538-4357/aa6e57.
Ullrich, P. A., T. Melvin, C. Jablonowski, and A. Staniforth, 2014: A proposed baroclinic wave test case for deep- and shallow-atmosphere dynamical cores. Quart. J. Roy. Meteor. Soc., 140, 1590–1602, https://doi.org/10.1002/qj.2241.
Walters, D. N., and Coauthors, 2017: The Met Office Unified Model global atmosphere 6.0/6.1 and JULES global land 6.0/6.1 configurations. Geosci. Model Dev., 10, 1487–1520, https://doi.org/10.5194/gmd-10-1487-2017.
White, A. A., and R. A. Bromley, 1995: Dynamically consistent, quasi-hydrostatic equations for global models with a complete representation of the Coriolis force. Quart. J. Roy. Meteor. Soc., 121, 399–418, https://doi.org/10.1002/qj.49712152208.
White, A. A., and A. Staniforth, 2008: A generalized thermal wind equation and some non-separable exact solutions of the flow equations for three-dimensional spherical atmospheres. Quart. J. Roy. Meteor. Soc., 134, 1931–1939, https://doi.org/10.1002/qj.323.
White, A. A., and N. Wood, 2012: Consistent approximate models of the global atmosphere in non-spherical geopotential coordinates. Quart. J. Roy. Meteor. Soc., 138, 980–988, https://doi.org/10.1002/qj.972.
White, A. A., B. J. Hoskins, I. Roulstone, and A. Staniforth, 2005: Consistent approximate models of the global atmosphere: Shallow, deep, hydrostatic, quasi-hydrostatic and non-hydrostatic. Quart. J. Roy. Meteor. Soc., 131, 2081–2107, https://doi.org/10.1256/qj.04.49.
Wood, N., and Coauthors, 2014: An inherently mass-conserving semi-implicit semi-Lagrangian discretization of the deep-atmosphere global non-hydrostatic equations. Quart. J. Roy. Meteor. Soc., 140, 1505–1520, https://doi.org/10.1002/qj.2235.
Zaron, E. D., B. S. Chua, P. A. Reinecke, J. Michalakes, J. D. Doyle, and L. Xu, 2022: The tangent-linear and adjoint models of the NEPTUNE dynamical core. Tellus, 74A, 399–411, https://doi.org/10.16993/tellusa.146.
Zawdie, K. A., M. S. Dhadly, S. E. McDonald, F. Sassi, C. Coker, and D. P. Drob, 2020: Day-to-day variability of the bottomside ionosphere. J. Atmos. Sol.-Terr. Phys., 205, 105299, https://doi.org/10.1016/j.jastp.2020.105299.