Vertical Coordinates and Upper Boundary Conditions. When selecting a vertical coordinate, there are three primary considerations to keep in mind:

Similar documents
Lateral Boundary Conditions

Synoptic Meteorology II: Potential Vorticity Inversion and Anomaly Structure April 2015

MODEL TYPE (Adapted from COMET online NWP modules) 1. Introduction

Lecture 23. Vertical coordinates General vertical coordinate

Synoptic Meteorology I: Skew-T Diagrams and Thermodynamic Properties

Grid-Based Methods for Numerical Weather Prediction

Split explicit methods

(Adapted from COMET online NWP modules) 1. Introduction

Final Examination, MEA 443 Fall 2008, Lackmann

Vertical structure. To conclude, we will review the critical factors invloved in the development of extratropical storms.

A semi-implicit non-hydrostatic covariant dynamical kernel using spectral representation in the horizontal and a height based vertical coordinate

Model Initialization

Mesoscale predictability under various synoptic regimes

Synoptic Meteorology I: Lab 1 Discussion. 25 September From page 14 of Meteorological Data and an Introduction to Synoptic Analysis,

Fronts in November 1998 Storm

Synoptic Meteorology II: Self-Development in the IPV Framework. 5-7 May 2015

Lower-Tropospheric Height Tendencies Associated with the Shearwise and Transverse Components of Quasigeostrophic Vertical Motion

Synoptic Meteorology I: Other Force Balances

Mesoscale Predictability of Terrain Induced Flows

Synoptic Meteorology I: Isoplething Example. From page 15 of Meteorological Data and an Introduction to Synoptic Analysis,

The Advanced Research WRF (ARW) Dynamics Solver

dv dt = f (M M g ) (1)

CHAPTER 4. THE HADLEY CIRCULATION 59 smaller than that in midlatitudes. This is illustrated in Fig. 4.2 which shows the departures from zonal symmetry

The dynamics of high and low pressure systems

7) Numerical Weather Prediction

Weather report 28 November 2017 Campinas/SP

Chapter 1. Governing Equations of GFD. 1.1 Mass continuity

Description of. Jimy Dudhia Dave Gill. Bill Skamarock. WPS d1 output. WPS d2 output Real and WRF. real.exe General Functions.

Dynamical coupling between the middle atmosphere and lower thermosphere

Errata: Midlatitude Synoptic Meteorology (July, 2014)

INTERPRETATION GUIDE TO MSG WATER VAPOUR CHANNELS

Numerical Weather Prediction. Meteorology 311 Fall 2010

The Shallow Water Equations

Thermodynamic Energy Equation

A mechanistic model study of quasi-stationary wave reflection. D.A. Ortland T.J. Dunkerton NorthWest Research Associates Bellevue WA

Chapter 3. Stability theory for zonal flows :formulation

P1.16 ADIABATIC LAPSE RATES IN TORNADIC ENVIRONMENTS

Synoptic Meteorology II: Frontogenesis Examples Figure 1

Isentropic Analysis. Much of this presentation is due to Jim Moore, SLU

Calculation and Application of MOPITT Averaging Kernels

196 7 atmospheric oscillations:

1/18/2011. Conservation of Momentum Conservation of Mass Conservation of Energy Scaling Analysis ESS227 Prof. Jin-Yi Yu

USING PROGRAM HURRICANE. Download all the files, including the (empty) output subdirectory into a new folder on your machine.

Mesoscale Atmospheric Systems. Surface fronts and frontogenesis. 06 March 2018 Heini Wernli. 06 March 2018 H. Wernli 1

Conservation of Mass Conservation of Energy Scaling Analysis. ESS227 Prof. Jin-Yi Yu

where p oo is a reference level constant pressure (often 10 5 Pa). Since θ is conserved for adiabatic motions, a prognostic temperature equation is:

Introduction to Isentropic Coordinates: a new view of mean meridional & eddy circulations. Cristiana Stan

Weather Research and Forecasting Model. Melissa Goering Glen Sampson ATMO 595E November 18, 2004

Typhoon Relocation in CWB WRF

Quasi-equilibrium Theory of Small Perturbations to Radiative- Convective Equilibrium States

The dynamics of a simple troposphere-stratosphere model

NWP Equations (Adapted from UCAR/COMET Online Modules)

Isentropic Analysis. We can look at weather data in other ways besides on constantpressure. Meteorology 411 Iowa State University Week 11 Bill Gallus

Impacts of Turbulence on Hurricane Intensity

Supplement to the. Final Report on the Project TRACHT-MODEL. Transport, Chemistry and Distribution of Trace Gases in the Tropopause Region: Model

Idealized Nonhydrostatic Supercell Simulations in the Global MPAS

APPENDIX B. The primitive equations

Numerical Weather Prediction. Meteorology 311 Fall 2016

p = ρrt p = ρr d = T( q v ) dp dz = ρg

WRF Model Simulated Proxy Datasets Used for GOES-R Research Activities

Mesoscale Meteorology Assignment #3 Q-G Theory Exercise. Due 23 February 2017

Occlusion cyclogenesis II

and 24 mm, hPa lapse rates between 3 and 4 K km 1, lifted index values

Hydrostatic Equation and Thermal Wind. Meteorology 411 Iowa State University Week 5 Bill Gallus

A stable treatment of conservative thermodynamic variables for semi-implicit semi-lagrangian dynamical cores

High initial time sensitivity of medium range forecasting observed for a stratospheric sudden warming

Mixing and Turbulence

MEA 716 Exercise, BMJ CP Scheme With acknowledgements to B. Rozumalski, M. Baldwin, and J. Kain Optional Review Assignment, distributed Th 2/18/2016

Using hybrid isentropic-sigma vertical coordinates in nonhydrostatic atmospheric modeling

By convention, C > 0 for counterclockwise flow, hence the contour must be counterclockwise.

For the operational forecaster one important precondition for the diagnosis and prediction of

INTRODUCTION TO METEOROLOGY PART ONE SC 213 MAY 21, 2014 JOHN BUSH

Four ways of inferring the MMC. 1. direct measurement of [v] 2. vorticity balance. 3. total energy balance

ABSTRACT 2 DATA 1 INTRODUCTION

The WRF NMM Core. Zavisa Janjic Talk modified and presented by Matthew Pyle

6 Two-layer shallow water theory.

AMS 17th Conference on Numerical Weather Predition, 1-5 August 2005, Washington D.C. Paper 16A.3

1/18/2011. From the hydrostatic equation, it is clear that a single. pressure and height in each vertical column of the atmosphere.

PRACTICAL INTERPRETATION OF RASP SOUNDINGS. Jean Oberson, February 2010.

P2.2 REDUCING THE IMPACT OF NOISE ABATEMENT PRACTICES ON AIRPORT CAPACITY BY FORECASTING SITUATIONAL DEPENDENT AIRCRAFT NOISE PROPAGATION

General Curvilinear Ocean Model (GCOM): Enabling Thermodynamics

GFD 2013 Lecture 4: Shallow Water Theory

Dynamic Inference of Background Error Correlation between Surface Skin and Air Temperature

Earth and Space Science. Teacher s Guide

4. Atmospheric transport. Daniel J. Jacob, Atmospheric Chemistry, Harvard University, Spring 2017

Atmospheric dynamics and meteorology

t tendency advection convergence twisting baroclinicity

P1.1 THE QUALITY OF HORIZONTAL ADVECTIVE TENDENCIES IN ATMOSPHERIC MODELS FOR THE 3 RD GABLS SCM INTERCOMPARISON CASE

1.4 CONNECTIONS BETWEEN PV INTRUSIONS AND TROPICAL CONVECTION. Beatriz M. Funatsu* and Darryn Waugh The Johns Hopkins University, Baltimore, MD

Daniel J. Jacob, Models of Atmospheric Transport and Chemistry, 2007.

9D.3 THE INFLUENCE OF VERTICAL WIND SHEAR ON DEEP CONVECTION IN THE TROPICS

Vertical discretisation for advanced sounder fast radiative transfer models:

Effects of thermal tides on the Venus atmospheric superrotation

Model description of AGCM5 of GFD-Dennou-Club edition. SWAMP project, GFD-Dennou-Club

29th Monitoring Research Review: Ground-Based Nuclear Explosion Monitoring Technologies

Upgrade of JMA s Typhoon Ensemble Prediction System

1 of 8 5/23/03 9:43 AM

Project 3 Convection and Atmospheric Thermodynamics

Fundamentals of Atmospheric Radiation and its Parameterization

Fluctuation dynamo amplified by intermittent shear bursts

Transcription:

Vertical Coordinates and Upper Boundary Conditions Introduction to Vertical Coordinate Systems Numerical models can be formulated with one of many vertical coordinates. A given numerical model will typically use transformed versions of the primitive equations applicable on a vertical coordinate surface chosen by the model developer. Thus, users of a given numerical model do not have the responsibility of choosing the vertical coordinate; their ability to choose the vertical coordinate rests solely in their choice of numerical model. When selecting a vertical coordinate, there are three primary considerations to keep in mind: Does the chosen vertical coordinate system permit unevenly distributed vertical levels? Strictly speaking, this consideration lies with the numerical model itself and not with the choice of vertical coordinate. Nearly all numerical models used for real-data simulations allow for unevenly distributed vertical levels. By contrast, some numerical models used for idealized simulations only permit a fixed distribution with constant spacing between vertical levels. Generally speaking, it is preferable that variable vertical resolution be permitted, such that more levels can be placed at the altitudes where sharp vertical gradients in meteorological quantities typically exist without having to add more vertical levels throughout the model atmosphere. Sharp vertical gradients often exist in the planetary boundary layer, such as with capping inversions, but also may occur near the tropopause and in association with fronts and jets. What is the relationship between the chosen vertical coordinate system and terrain? In numerical models that utilize uneven terrain, whether the terrain is meant to represent the actual Earth or an idealized topographic surface, it is possible for the vertical surfaces of selected vertical coordinate systems to intersect the terrain. This poses a computational challenge that must be addressed in some fashion so as to ensure robust results. Generally speaking, it is preferable that vertical coordinate surfaces do not intersect the terrain; e.g., that they be terrain-following in some fashion. Upon transforming the horizontal pressure gradient terms in the momentum equations to the chosen vertical coordinate system, do one or two terms result? While it is fairly straightforward to transform the horizontal pressure gradient terms in the momentum equations between chosen vertical coordinate systems, the transformation itself can result in an added term for the horizontal pressure gradient. For example, when we introduced the WRF-ARW model equations, we transformed the horizontal pressure gradient term in the u-momentum equation from constant height surfaces to the terrainfollowing η vertical coordinate used by the WRF-ARW model: Vertical Coordinate Systems and Upper Boundaries, Page 1

1 p x z (constant height) p p d x d x (terrain-following, coupled to the dry air mass) As finite difference approximations may be used to solve the primitive equations, a second horizontal pressure gradient term introduces a potential additional source of error into the model. This is particularly troublesome because truncation errors in the computation of the horizontal pressure gradient can be of similar magnitude to horizontal variability in the horizontal pressure gradient itself. Modern models such as the WRF-ARW model typically address this by recasting pressure, geopotential height, and dry air mass into perturbation form, such that truncation errors of finite differences calculated using smaller values are themselves smaller. However, generally speaking, it remains desirable to have just a single horizontal pressure gradient term. We now wish to consider the strengths and weaknesses of several vertical coordinate systems in light of the aforementioned considerations. Height Above Sea Level Constant Height (z) Coordinate This coordinate is perhaps the most intuitive: sea level is at a height z = 0, and the height z increases with increasing distance above sea level. It is the vertical coordinate that we are first introduced to when considering the primitive equations as undergraduate students. The horizontal pressure gradient term in this vertical coordinate is of the form: 1 z p where the subscript of z on the gradient operator indicates that it is applied on a constant height surface. There is only one horizontal pressure gradient term in the u- and v-momentum equations when formulated on constant height surfaces. Thus, our third criterion above is met. This vertical coordinate system poses no inhibition against having unevenly distributed vertical levels, such that it meets our first criterion above so long as the model itself permits such a distribution of vertical levels. An example of unevenly distributed constant height surfaces relative to a hypothetical sloping terrain surface is given in Figure 1. This figure, however, also highlights the primary shortcoming of the constant height vertical coordinate: except for flat terrain at sea level, it is a terrain-intersecting vertical coordinate. Vertical Coordinate Systems and Upper Boundaries, Page 2

Figure 1. Hypothetical depiction of unevenly distributed constant height (z) surfaces relative to an idealized sloping terrain profile. Consider partial derivatives in both the horizontal and vertical directions evaluated adjacent to the terrain along a given constant height surface such as the z = 2000 m surface in Figure 1. Evaluating these partial derivatives involves one or more points where atmospheric quantities are undefined due to being below ground. Furthermore, the discontinuous nature of atmospheric fields adjacent to the terrain along constant height surfaces renders impossible the evaluation of horizontal partial derivatives using spectral methods with this vertical coordinate. While there are methods to address the computation of partial derivatives along constant height surfaces near sloping terrain, as described within the course text, these can be computationally expensive. Alternatively, one can use forward or backward finite differences to compute partial derivatives adjacent to sloping terrain, but these are associated with an unacceptably high amount of truncation error. Furthermore, each of these alternatives require different methods to compute the partial derivatives adjacent to terrain from those in the free atmosphere. Consequently, they pose a technical challenge to implement within a numerical model. As a result of this shortcoming, the constant height vertical coordinate is typically not used within modern models. The primary exception to this lies with models utilized exclusively for idealized simulations that do not permit uneven terrain. Vertical Coordinate Systems and Upper Boundaries, Page 3

Pressure Isobaric (p) Surfaces We are also likely quite familiar with the primitive equations cast into the isobaric, or constant pressure, vertical coordinate. Pressure, a function of the weight of the air above you, is highest at the surface and decreases log-linearly with increasing height away from the surface. The transformation of the primitive equations into the isobaric coordinate system results in only one term for the horizontal pressure gradient for each of the horizontal momentum equations. This term takes the form, where the geopotential Φ = gz and the subscript of p on the gradient p operator indicates that it is applied on an isobaric surface. As a result, our third criterion is met, just as it was with the z vertical coordinate. As did the constant height vertical coordinate, the isobaric coordinate system poses no inhibition against having unevenly distributed vertical levels, such that it meets our first criterion above so long as the model itself permits such a distribution of vertical levels. An example of unevenly distributed isobaric surfaces relative to a hypothetical sloping terrain surface is given in Figure 2. This figure, however, also highlights the primary shortcoming of the isobaric vertical coordinate: it is also a terrain-intersecting vertical coordinate. Figure 2. Hypothetical depiction of unevenly distributed isobaric surfaces relative to an idealized sloping terrain profile. The slope of the isobaric surfaces depicted here is a function of the prevailing meteorology, indicating relatively high (low) pressure to the left/west (right/east). The same challenges with computing partial derivatives adjacent to sloping terrain, whether using finite difference or spectral methods, noted for the z coordinate hold for the isobaric vertical coordinate. The isobaric vertical coordinate poses an additional challenge, however, in that the isobaric surfaces themselves change altitude with time as a function of the prevailing meteorology. In other words, at a given horizontal grid point, a given isobaric surface may be above ground at Vertical Coordinate Systems and Upper Boundaries, Page 4

one time but below ground at another time. As a result, the isobaric vertical coordinate is typically not used within numerical models. Potential Temperature Isentropic (θ) Surfaces As an undergraduate student, you hopefully were introduced to potential temperature, which is a function of both temperature T and pressure p as expressed by Poisson s equation: p0 T p where p0 is a reference-state pressure and is typically equal to 1000 hpa. Typically, though not always, potential temperature increases with increasing height. Under dry adiabatic conditions, potential temperature is conserved following the motion. This makes potential temperature an appealing vertical coordinate: flow across isentropes, which constitutes vertical motion in the isentropic vertical coordinate system, is small and only occurs physically when diabatic processes are active. If vertical motion is small, vertical advection and thus numerical diffusion in the vertical are small, which is also appealing. A further benefit of the isentropic vertical coordinate is that vertical levels i.e., isentropic surfaces are naturally packed where temperature gradients are large, such as with fronts. A representative example is given in Figure 3. At left, a vertical cross-section from San Diego, CA to Medford, OR in isobaric coordinates is depicted. The cold front is located between 700-600 hpa to the south and slopes upward toward the north, intersecting the tropopause between 400-300 hpa near Oakland, CA. Isentropes are tightly-packed with the front. By thermal wind balance, it should not be surprising that there is strong vertical wind shear with the front, particularly between Point Arguello, CA and Oakland, CA. Thus, there exist sharp horizontal and vertical gradients in both potential temperature and wind across the front. At right, the same data are plotted, except using an isentropic vertical coordinate system. The cold front is found between the 300-320 K isentropic surfaces, as expected from the sloping isentropes at left. For a chosen vertical level distribution, this inherently allows higher vertical resolution where the sharpest gradients exist. The same is true across any temperature inversion, and thus our third criterion is met without the need to explicitly specify unevenly distributed vertical levels. Both horizontal and vertical wind gradients are reduced, although the coordinate transformation does introduce relatively sharp horizontal gradients in height, pressure, and temperature (not shown). R d c p Vertical Coordinate Systems and Upper Boundaries, Page 5

Figure 3. Vertical cross-sections (south-north) of a cold front in (left) isobaric and (right) isentropic coordinates. In both panels, isentropes (every 5 K at left, every 10 K at right) are given by the solid lines and isotachs (every 10 m s -1 ) are given by the dashed lines. The shaded grey area in the center of each panel spans the same physical volume of atmosphere encompassed by the cold front between Point Arguello, CA and Oakland, CA. Figure reproduced from Benjamin (1989, Mon. Wea. Rev.), their Figure 1. There exist three primary shortcomings of the isentropic vertical coordinate, however, that limit its use for numerical weather prediction: Where the lapse rate exceeds the dry adiabatic lapse rate, such as in Figure 4, a given isentropic surface is non-uniquely identified with a given altitude (and thus meteorological conditions). This most commonly occurs in the case of strongly-heated terrain during the local daytime hours. While a correction term can be added to the model so as to prevent the lapse rate from becoming superadiabatic, such a term is non-physical and may result in inaccurate model results near the surface. As with isobaric surfaces, isentropic surfaces can intersect the ground, and whether a given grid point is above or below ground can change as a function of the meteorology. This latter issue is arguably more common with isentropic surfaces than with isobaric surfaces given the former s comparatively stronger ties to temperature. The transformation of the horizontal pressure gradient term into isentropic coordinates results in the addition of a second term. Specifically, the horizontal pressure gradient term applicable on isentropic surfaces takes the form: M gz c p T Vertical Coordinate Systems and Upper Boundaries, Page 6

where M is the Montgomery streamfunction. The subscript of θ on the gradient operator indicates that it is applied on isentropic surfaces. As there may exist sharp horizontal gradients in both z and T on an isentropic surface such as associated with a front relatively large truncation errors may result when finite difference methods are used. As these errors may not necessarily cancel, this results in a potentially large source of error. As a result of these shortcomings, the use of isentropic vertical coordinates is rare in modern NWP. Figure 4. (left) 0000 UTC 3 July 2012 skew T-ln p diagram for Norman, OK. The temperature (dewpoint temperature) trace is given by the rightmost (leftmost) thick black line. (right) Tabular display of data from below 913.2 hpa from the 0000 UTC 3 July 2012 Norman, OK skew T-ln p diagram. Note the superadiabatic lapse rate at the surface at left and the corresponding decrease in potential temperature (THTA) between 973 hpa and 964 hpa at right. Figure and data obtained from the University of Wyoming. Terrain-Following Vertical Coordinates As we discovered with the WRF-ARW model, it is possible to utilize a terrain-following vertical coordinate. The primary advantage of doing so is to eliminate the possibility for vertical coordinate surfaces to intersect the terrain, which was a major shortcoming with the constant height, isobaric, and isentropic vertical coordinates noted above. A generalized terrain-following vertical coordinate σ may be formulated with respect to either pressure p (left) or height z (right): Vertical Coordinate Systems and Upper Boundaries, Page 7

p p p s t p t z z t t z z s For both pressure- and height-based formulations, σ = 0 at the top of the model and σ = 1 at the terrain surface. The terrain influence on coordinate surface slope decreases linearly with increasing altitude. Slightly modifications can allow coordinate surfaces to transition from terrain-following to isobaric at altitudes well above the terrain; see Klemp (2011, Mon. Wea. Rev.) for more details. Note that p need not be the full pressure but can be some derivative thereof, such as dry hydrostatic pressure as in the WRF-ARW model. Pressure pt or height zt at the top of the model is specified by the user. The surface height zs is fixed to the terrain, whereas the surface pressure ps can change by a small amount through the simulation as a function of the prevailing meteorology. Because surface pressure can change through a model simulation, there exists a slight preference to height-based rather than pressure-based terrain-following vertical coordinates. However, most modern models use pressure-based terrain-following vertical coordinates, and practically all use some form of a terrain-following vertical coordinate. Representative examples of pressure-based vertical coordinates include the WRF-ARW, RAP/HRRR, GFS, NAM, and ECMWF models, and representative examples of height-based vertical coordinates include the CM1 and MPAS models. Models that use terrain-following vertical coordinates typically allow non-uniform vertical level distribution, so that our first consideration related to the choice of vertical coordinate is met. Figure 5 depicts an example of pressure-based terrain-following vertical coordinate surfaces, distributed between the terrain surface and the model top. The terrain-following coordinate surfaces means that our second consideration related to the choice of vertical coordinate is also met. Figure 5. Cross-section of terrain-following pressure-based vertical coordinate surfaces for pt = 500 hpa. Figure reproduced from Warner et al. (1978, Mon. Wea. Rev.), their Figure 9. Vertical Coordinate Systems and Upper Boundaries, Page 8

This only leaves the computation of the horizontal pressure gradient, our third consideration. The generalized form of the transform operator between constant height z and terrain-following σ surfaces takes the form: z z where subscripts of z and σ indicate the coordinate surface on which the quantity is evaluated. The original and transformed versions of the horizontal pressure gradient term in the u-momentum equation for the terrain-following pressure-based vertical coordinate used by the WRF-ARW model were given above and are reproduced below for convenience: z 1 p x z (constant height) p p d x d x (terrain-following, coupled to the dry air mass) After the basic conversion to the terrain-following vertical coordinate, further manipulation of the second term was performed to arrive at the form depicted above at right. Similar results would be obtained if a height-based terrain-following vertical coordinate were utilized or if a different model s formulation of the governing equations were considered. Horizontal changes in pressure p and geopotential Φ along the terrain-following coordinate surface are particularly large in the presence of sharp horizontal gradients in terrain height. This can be understood given the definitions of pressure (related to the weight of the air above you) and the geopotential (related to the geometric height z). Consequently, the horizontal pressure gradient can be comparatively large in regions of sloped terrain, such that truncation error resulting from the computation of the two horizontal pressure gradient terms may also be large. As noted above, this is typically addressed by recasting the relevant terms into perturbation form to reduce truncation error magnitude. It is also possible to devise a terrain-following vertical coordinate that is a function of potential temperature. Generally speaking, such vertical coordinates are terrain-following near the surface and transition to isentropic surfaces at some specified distance above the terrain. However, with a few exceptions the now-replaced RUC model being one notable exception terrain-following isentropic-based vertical coordinates are not used in modern models. Vertical Coordinate Systems and Upper Boundaries, Page 9

Upper Boundary Conditions Numerical models generally do not simulate the full depth of the atmosphere; i.e., from the surface to where p = 0 hpa. Rather, the depth of the model atmosphere is truncated, usually to an isobaric surface somewhere in the stratosphere. However, vertically-propagating internal gravity waves, such as may be generated by intense thunderstorm updrafts impinging upon the tropopause or by flow passing over sloped terrain, can propagate over a large vertical distance and potentially reach the model top. What, then, should happen in the model atmosphere as these internal gravity waves reach its uppermost limit? Most modern models use a rigid lid to represent the boundary at the top of the model atmosphere. This requires specifying an appropriate upper boundary condition. In the WRF-ARW model, this upper boundary condition is specified by linearly interpolating realistic profiles of model variables from the top of the atmosphere to the top of the model atmosphere. This upper boundary condition works reasonably well when the top of the model is at altitudes where meteorological variability is small (e.g., near the stratopause). It doesn t work as well at lower altitudes where meteorological variability is large. See Cavallo et al. (2011, Mon. Wea. Rev.) for more details regarding this upper boundary condition, noting that other formulations for upper boundary conditions are possible. To mitigate the non-physical reflection of vertically-propagating waves as they impinge upon the rigid lid, an absorbing or damping layer can be used within the upper reaches of the model. There are many forms for this layer, with the two most common being viscous and Rayleigh damping. Viscous damping involves defining an absorbing layer in the upper reaches of the model. In this layer, the eddy viscosity coefficients Kh (horizontal) and Kv (vertical) have default values at the bottom of the layer that increase to a specified maximum value at the top of the layer. Eddy viscosity coefficients act to reduce horizontal and/or vertical gradients of meteorological fields where they exist within the model atmosphere. The influence of a viscous damping layer of 20 km depth in model with a model top at z = 50 km is depicted in Figure 6. The viscous damping option available within the WRF-ARW model takes the form: K dh x t 2 ztop z g cos (horizontal) 2 zd K dv z t 2 ztop z g cos (vertical) 2 zd In the above, the Kdh and Kdv are the eddy viscosity coefficients applicable in the damping layer; γg is a user-specified nondimensional damping coefficient, ztop is the height of the model top, and zd is the depth of the damping layer. The recommended values for γg range between 0.01 and 0.1 Vertical Coordinate Systems and Upper Boundaries, Page 10

and the default value for zd is 5 km. At the top of the model, where z = ztop, the cos functions equal 1. At the bottom of the damping layer, where ztop z = zd, the cos functions equal 0. Figure 6. Vertical motion (contours; negative values shaded) from 2-D model simulations of flow over elevated topography (dark shading at bottom) for simulations with (a) a viscous damping layer of 20 km depth at the top of the model and (b) no viscous damping. Figure reproduced from Warner (2011), their Figure 3.50. Rayleigh damping also involves defining an absorbing layer in the upper portions of the model. Here, however, model variables are relaxed (or nudged) toward a pre-defined reference state. For any model variable α, a generic formulation for a Rayleigh damping operator is of the form: t z Here, the time tendency of α is a function of the damping operator τ, which is a function of height z, and the departure of the value of α from its reference state value. In the WRF-ARW model, a form of this Rayleigh damping is available, where: u t v zu u zv v t z t t z Reference state fields are a function of z only. The reference state vertical velocity is assumed to be zero. Because the model vertical coordinate is not a function of height, the height of the vertical coordinate surfaces changes with time. Thus, the reference state values on model coordinate surfaces must be adjusted accordingly at each time step. The Rayleigh damping function takes the form: Vertical Coordinate Systems and Upper Boundaries, Page 11

z z d sin 2 1 2 z zd where the Rayleigh damping is applied only over the damping layer depth specified by zd. The γd is a user-specified damping coefficient. The recommended value for γd is approximately equal to 0.003 s -1. At the top of the model, where z = ztop, the sin 2 function is equal to 1 and damping takes its maximum value of top r. At the bottom of the damping layer, where ztop z = zd, the sin 2 function is equal to 0 and damping is not applied. It is also possible to apply an implicit form of Rayleigh damping only to the vertical velocity. In this formulation, the Rayleigh damping is applied to the perturbation vertical velocity within the acoustic time step loop of the model s time integration and influences both the coupled vertical velocity W and, given that W appears in its definition, the geopotential Φ. This Rayleigh damping takes the form: ~ W' ' W'' W '' Here, τ takes the same form as for the traditional Rayleigh damping, albeit with a recommended value of γd of approximately 0.2 s -1 ~. W '' is the value of W'' at the end of the acoustic time step while W'' is the value prior to the acoustic time step. This formulation is akin to including an implicit damping term within the vertical momentum equation and adding implicit vertical diffusion to the vertical velocity. Klemp et al. (2008, Mon. Wea. Rev.) should be consulted for more information regarding this formulation. Alternatively, the model top may be represented by a free surface, wherein the model atmosphere is treated as a fluid that lies beneath a second fluid above the top of the model atmosphere. There exists no flow across a free surface, although the height of the free surface itself may change. Despite this, free surface boundaries also reflect vertically-propagating waves, necessitating the use of an absorbing or damping layer in the upper portions of the simulation domain. Finally, the model top may instead be represented by a radiative boundary condition, wherein the energy associated with vertically-propagating waves is permitted to be radiated upward and out of the simulation domain. In models that utilize a radiative upper boundary condition, no absorbing or damping layer is utilized, and the top of the model is fixed at some user-specified altitude. See also Klemp and Durran (1983, Mon. Wea. Rev.) for more details regarding radiative boundary condition formulation and their performance under a wide range of conditions for idealized terrain. z Vertical Coordinate Systems and Upper Boundaries, Page 12