assessment of physical parameterizations using a global climate

16
Assessment of Physical Parameterizations Using a Global Climate Model with Stretchable Grid and Nudging O. COINDREAU Commisariat à l’Energie Atomique, Bruyères-le-Châtel, France F. HOURDIN, M. HAEFFELIN, A. MATHIEU, AND C. RIO Laboratoire de Météorologie Dynamique, CNRS, IPSL, Paris, France (Manuscript received 19 October 2005, in final form 16 May 2006) ABSTRACT The Laboratoire de Météorologie Dynamique atmospheric general circulation model with zooming ca- pability (LMDZ) has been used in a nudged mode to enable comparison of model outputs with routine observations and evaluate the model physical parameterizations. Simulations have been conducted with a stretched grid refined over the vicinity of Paris, France, where observations, collected at the Trappes station (Météo-France) and at the Site Instrumental de Recherche par Télédétection Atmosphérique observatory, are available. For the purpose of evaluation of physical parameterizations, the large-scale component of the modeled circulation is adjusted toward ECMWF analyses outside the zoomed area only, whereas the inside region can evolve freely. A series of sensitivity experiments have been performed with different param- eterizations of land surface and boundary layer processes. Compared with previous versions of the LMDZ model, a “thermal plume model,” in association with a constant resistance to evaporation improves agree- ment with observations. The new parameterization significantly improves the representation of seasonal and diurnal cycles of near-surface meteorology, the day-to-day variability of planetary boundary layer height, and the cloud radiative forcing. This study emphasizes the potential of using a climate model with a nudging and zooming capability to assess model physical parameterizations. 1. Introduction The skill of atmospheric general circulation models (GCMs) relies, to a large part, on the representation of subgrid-scale atmospheric processes such as turbulent mixing in the planetary boundary layer (PBL), cumulus and dry convection, and surface thermodynamics. Those processes can only be represented statistically in the models through so-called parameterizations. In the frame of climate change projections, those parameter- izations must, as much as possible, rely on physics and not be tied to reproducing current observations. Im- provement and validation of those parameterizations is a constant effort in climate modeling groups. Different strategies have been developed. In the first approach, model changes are directly assessed in the full 3D cli- mate model by comparing simulations and observations (or meteorological analysis) in terms of statistics (means, standard deviations, correlations, etc.). This approach is at the basis of intercomparison programs such as the Atmospheric Model Intercomparison Proj- ect (Gates 1992). This approach is important because one validates at once the whole set of physical param- eterizations coupled to large-scale dynamics. The dis- advantage is that it is often difficult to separate which of the parameterizations (or combination of them) is re- sponsible for a discrepancy between observations and model. To overcome this difficulty, a second approach was developed in the last two decades that consists of com- paring single-column versions of the set (or a subset) of physical parameterizations with results of explicit simu- lations of the same phenomenon, using the so-called cloud resolving models (Guichard et al. 2004) or large eddy simulations (LESs; Ayotte et al. 1996; Lenderink et al. 2004). This approach contributed to improving the description of some phenomenon and has led to impor- tant international programs such as the European Corresponding author address: Dr. Olivia Coindreau, CEA, DASE/LDG/SEG, BP 12, 91 680 Bruyéres-le-Châtel, France. E-mail: [email protected] 1474 MONTHLY WEATHER REVIEW VOLUME 135 DOI: 10.1175/MWR3338.1 © 2007 American Meteorological Society MWR3338

Upload: vanxuyen

Post on 05-Jan-2017

214 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Assessment of Physical Parameterizations Using a Global Climate

Assessment of Physical Parameterizations Using a Global Climate Model withStretchable Grid and Nudging

O. COINDREAU

Commisariat à l’Energie Atomique, Bruyères-le-Châtel, France

F. HOURDIN, M. HAEFFELIN, A. MATHIEU, AND C. RIO

Laboratoire de Météorologie Dynamique, CNRS, IPSL, Paris, France

(Manuscript received 19 October 2005, in final form 16 May 2006)

ABSTRACT

The Laboratoire de Météorologie Dynamique atmospheric general circulation model with zooming ca-pability (LMDZ) has been used in a nudged mode to enable comparison of model outputs with routineobservations and evaluate the model physical parameterizations. Simulations have been conducted with astretched grid refined over the vicinity of Paris, France, where observations, collected at the Trappes station(Météo-France) and at the Site Instrumental de Recherche par Télédétection Atmosphérique observatory,are available. For the purpose of evaluation of physical parameterizations, the large-scale component of themodeled circulation is adjusted toward ECMWF analyses outside the zoomed area only, whereas the insideregion can evolve freely. A series of sensitivity experiments have been performed with different param-eterizations of land surface and boundary layer processes. Compared with previous versions of the LMDZmodel, a “thermal plume model,” in association with a constant resistance to evaporation improves agree-ment with observations. The new parameterization significantly improves the representation of seasonaland diurnal cycles of near-surface meteorology, the day-to-day variability of planetary boundary layerheight, and the cloud radiative forcing. This study emphasizes the potential of using a climate model witha nudging and zooming capability to assess model physical parameterizations.

1. Introduction

The skill of atmospheric general circulation models(GCMs) relies, to a large part, on the representation ofsubgrid-scale atmospheric processes such as turbulentmixing in the planetary boundary layer (PBL), cumulusand dry convection, and surface thermodynamics.Those processes can only be represented statistically inthe models through so-called parameterizations. In theframe of climate change projections, those parameter-izations must, as much as possible, rely on physics andnot be tied to reproducing current observations. Im-provement and validation of those parameterizations isa constant effort in climate modeling groups. Differentstrategies have been developed. In the first approach,model changes are directly assessed in the full 3D cli-mate model by comparing simulations and observations

(or meteorological analysis) in terms of statistics(means, standard deviations, correlations, etc.). Thisapproach is at the basis of intercomparison programssuch as the Atmospheric Model Intercomparison Proj-ect (Gates 1992). This approach is important becauseone validates at once the whole set of physical param-eterizations coupled to large-scale dynamics. The dis-advantage is that it is often difficult to separate which ofthe parameterizations (or combination of them) is re-sponsible for a discrepancy between observations andmodel.

To overcome this difficulty, a second approach wasdeveloped in the last two decades that consists of com-paring single-column versions of the set (or a subset) ofphysical parameterizations with results of explicit simu-lations of the same phenomenon, using the so-calledcloud resolving models (Guichard et al. 2004) or largeeddy simulations (LESs; Ayotte et al. 1996; Lenderinket al. 2004). This approach contributed to improving thedescription of some phenomenon and has led to impor-tant international programs such as the European

Corresponding author address: Dr. Olivia Coindreau, CEA,DASE/LDG/SEG, BP 12, 91 680 Bruyéres-le-Châtel, France.E-mail: [email protected]

1474 M O N T H L Y W E A T H E R R E V I E W VOLUME 135

DOI: 10.1175/MWR3338.1

© 2007 American Meteorological Society

MWR3338

Page 2: Assessment of Physical Parameterizations Using a Global Climate

Cloud-Resolving Modeling Project, the EuropeanCloud Systems research project, and the Global Energyand Water Cycle Experiment Cloud System Study.With this approach, it is possible to extract a number ofdiagnostics from the explicit simulations. It is alsoeasier to identify which component of the parameter-ization is correct. The disadvantage here is that theexplicit simulations are quite computationally demand-ing and the number of cases limited.

A complementary approach consists of comparingcontinuous simulations of the local meteorology, rely-ing for the large-scale forcing on the operational me-teorological analyses, with routine observations collectedat advanced atmospheric profiling observatories. Forinstance, Morcrette (2002) assesses the European Cen-tre for Medium-Range Weather Forecasts (ECMWF)model cloudiness and surface radiation by direct com-parisons with data from the Atmospheric RadiationMeasurement (ARM) program (Ackerman and Stokes2003) over a period of 1 month. The meteorologicalanalysis can also be used to define the boundary con-ditions of a limited-area model. For instance, Guichardet al. (2003) make use of the ARM data for directlyevaluating the predictions and parameterizations of thefifth-generation Pennsylvania State University–NationalCenter for Atmospheric Research Mesoscale Model(MM5), and Chiriaco et al. (2006) use data from the SiteInstrumental de Recherche par Télédétection Atmo-sphérique (SIRTA) observatory (Haeffelin et al. 2005) toassess cloud physics parameterizations in the MM5. Ghanet al. (1999) compared and evaluated the prospects andlimitations of these different approaches in evaluating thephysical parameterizations. We present here a tool thatcan be classified in the third category. The Laboratoire deMétéorologie Dynamique GCM with zooming capability(LMDZ), an atmospheric GCM involved recently in theproduction of climate change simulations for the next In-tergovernmental Panel on Climate Change report (Martiet al. 2005), is used in a nudged mode. The model is forcedto stay close to the synoptic situation by relaxing thelarge-scale circulation toward the meteorological analy-ses. This nudging technique is well known in the commu-nity of transport modelers and can be presented as a poorversion of data assimilation (Jeuken et al. 1996). Here,real observations are replaced by the results of a formerassimilation procedure performed in meteorological cen-ters.

The first focus of this paper is to show the potentialof this approach for model development and validation.This is illustrated with a series of sensitivity experi-ments performed with the LMDZ model. In the presentstudy, a zooming capability of the GCM is used thatallows the grid to be refined around a site where rou-

tine observations are available. For the purpose ofevaluating the physical parameterizations, the large-scale component of the modeled circulation is adjustedtoward ECMWF analyses outside the zoomed areaonly, whereas the inside region can evolve freely. Theoutputs of the LMDZ model zoomed over the Parisarea are compared with observations of the SIRTA ob-servatory. It is shown in particular that despite nudging,model results are highly sensitive to the parameteriza-tion of boundary layer and surface processes.

The second focus of this paper is to test in a real 3Dcontext the “thermal plume model” (th) developed toaccount for nonlocal transport by mesoscale thermalcells in the convective boundary layer (Hourdin et al.2002). The influence of the PBL parameterization isevaluated by running the model in the configurationpreviously described and confronting model outputswith SIRTA observations. It allows the new thermalplume model to be compared with the standard PBLparameterization of the LMDZ model.

In section 2, a description is made of the model andits physical parameterizations. The nudging techniqueis also briefly described and results of the referenceexperiment are compared with ECMWF analyses. Ob-served meteorological data collected at the Trappes sta-tion (Météo-France), such as temperature and relativehumidity, are used to assess the sensitivity of the modelto the parameterization of the land surface (section 3)and boundary layer (section 4) schemes. Comparisonsare performed during the period January 2000–December 2003. Observations of the convective PBLrealized at SIRTA in May–June 2004 during an inten-sive observation period are compared in the last sectionwith the results obtained with the best set of physicalparameterizations.

2. Description of the model and observations

a. The LMDZ GCM

The GCM used in this study is based on a finite-difference formulation of the primitive equations ofmeteorology, first described by Sadourny and Laval(1984). The dynamical equations are discretized on alongitude–latitude Arakawa C-grid (Kasahara 1977)with zooming capability. Discretization in the vertical isdone by using a hybrid �–p coordinate system (Sim-mons and Burridge 1981) with 19 levels.

The physical package of the version used here is de-scribed in detail by Hourdin et al. (2006). The radiativetransfer model is a refined version of the scheme de-veloped by Fouquart and Bonnel (1980) for the solarpart, and the terrestrial part is based on Morcrette et al.

APRIL 2007 C O I N D R E A U E T A L . 1475

Page 3: Assessment of Physical Parameterizations Using a Global Climate

(1986). Turbulent mixing in the PBL is parameterizedby using a classical local closure with a turbulent coef-ficient that depends on the vertical shear of the hori-zontal wind and on the local Richardson number fol-lowing Laval et al. (1981). A countergradient term forpotential temperature is introduced to allow upwardheat transport in a neutral or slightly stable atmosphereas is frequently observed in the convective PBL (Dear-dorff 1972). Condensation is parameterized separatelyfor convective and nonconvective clouds. Moist convec-tion is accounted for with the Emanuel (1991) mass fluxscheme. Large-scale condensation and cloud cover forboth large-scale and convective clouds is prescribedthrough a probability distribution function of the sub-grid-scale total water following Bony and Emanuel(2001). The large-scale transport of vapor and con-densed water is computed with the finite-volumescheme of Van Leer (1977).

The standard parameterization scheme for land sur-face processes in LMDZ is the Schématisation desEchanges Hydriques à l’Interface Biosphère–Atmosphère (SECHIBA; Ducoudré et al. 1993; DeRosnay and Polcher 1998; Ducharne and Laval 2000)recently used in the dynamical vegetation model Orga-nizing Carbon and Hydrology in Dynamic Ecosystems(ORCHIDEE; Krinner et al. 2005). SECHIBA is a pa-rameterization of the hydrological exchanges betweenthe land surface and the atmosphere. There are twomoisture reservoirs of water in the soil: a superficialone, created as soon as precipitation is larger thanevaporation, and a lower one. As long as the superficialreservoir exists, the water content of the lower reser-voir can only increase by drainage between the two soillayers. When no upper reservoir exists, the deep oneworks as a simple bucket (Manabe 1969). The totaldepth of the soil is 2 m and its maximum water contentis 300 mm. The grid box is subdivided into several land

tiles, each corresponding to one hydrological soil col-umn. Water exchanges between columns are allowed inthe lower layer. Vegetation types are characterized bytheir root density distributions, their leaf area index,and their canopy resistance. In the region studied here(whose area is 120 � 120 km2), the land surface is com-posed of cultures (0.71), bare soil (0.24), deciduous for-est (0.04), and grassland (0.01). Thermal conduction be-low the surface is computed with an 11-layer modelfollowing Hourdin et al. (1993).

b. Model configuration

To assess physical parameterizations of the modelover a given area, the model is run with its zoomingcapability in a nudged mode. The region studied here isthe vicinity of Paris, France. At the center of the zoom,fixed at 48°N, 2°E, the resolution reaches about 120 km(see Fig. 1a). The small number of grid points (48 � 32)allows a series of pluriannual simulations to be per-formed even on a standard personal computer with amesh size that is typical of state-of-the-art climate mod-els.

Meteorological fields (wind, temperature, and hu-midity) are relaxed toward analyzed fields of theECMWF, by adding a nonphysical relaxation term tothe model equations:

�X

�t� F �X� �

Xa � X

�, �1�

where X represents any meteorological field, F is theoperator describing the dynamical and physical pro-cesses that determine the evolution of X, Xa is the ana-lyzed field of ECMWF, and � is the time constant. Thechoice of the right value of � is crucial as it determinesthe effect of the observations on the solution. As sug-gested by Hoke and Anthes (1976), the optimal value

FIG. 1. (a) LMDZ stretched grid with 48 � 32 points. (b) A zoom over Europe showing the nudging time constant. Contour levelsare 0.1, 0.3, 1, 3, and 9 days. The time constant is a maximum at the center of the zoomed area and decreases as the grid resolutiondecreases.

1476 M O N T H L Y W E A T H E R R E V I E W VOLUME 135

Page 4: Assessment of Physical Parameterizations Using a Global Climate

should depend on the observational accuracy, the kindof variable being adjusted, and the typical magnitude ofthe model forcing. In the LMDZ model, � can be cho-sen differently for each variable and each region.

Before applying relaxation, ECMWF data are inter-polated on the stretched grid of the LMDZ model. Thetime constant in Eq. (1) varies from 30 min outside thezoomed area to 10 days inside for temperature andwind fields (see Fig. 1b). Humidity fields are relaxedwith a larger time constant outside the zoomed area (5h). With these values of the relaxation coefficient, thelarge-scale component of the modeled circulation is ad-justed toward ECMWF analyses whereas the evolutionof meteorological fields in the zoomed area is condi-tioned by physical parameterizations of the model. Asshown in Fig. 2, simulated temperatures are close toanalyses away from the center of the zoomed area, in-dicating a strong forcing of the synoptic situation. Onthe contrary, simulated temperatures and analyses are

quite different at the center of the zoomed area, show-ing that a certain degree of freedom is left to param-eterizations.

A series of experiments, differing by their land sur-face and boundary layer schemes, are carried out forthe period January 1998–June 2004 [the last period,running from May to June 2004, corresponding to thewater vapor profiling intercomparison campaign(VAPIC) described below]. One year is needed toreach a regime state for the seasonal cycle of surfacehydrology (which is not initialized with observations).In the analysis presented hereafter, evaluations aremade on the period January 2000–June 2004, skippingthe first two years. The time step for large-scale dynam-ics is 1.5 min whereas most model physics are calculatedevery 30 min. The model variables we analyze are thoseof the grid box containing the point 48°N, 2°E.

c. Observations

Model variables such as temperature, relative humid-ity, soil moisture, evaporation, and precipitation rate,PBL height, and downward shortwave fluxes are inves-tigated and compared with measurements. Those mea-surements are listed in Table 1.

For comparisons of standard meteorological param-eters such as temperature, humidity, and precipitation,over long time periods, we used hourly data providedby the meteorological center of Trappes (48.8°N,2.0°E), France, operated by Météo-France. This sta-tion, located 30 km southwest of Paris, provided hourlyvalues of the standard parameters for the period Janu-ary 2000–December 2003.

For comparison of physical and radiative properties,we use the measurement synergies provided by theSIRTA observatory, located in Palaiseau (48.7°N,2.2°E), France, 25 km south of Paris. SIRTA gathers asuite of active (lidar and radar) and passive (radiom-eters) remote sensing instruments and in situ sensors todescribe atmospheric processes in the atmosphericboundary layer and the free troposphere (Haeffelin et

TABLE 1. List of available observations.

Location Period Sensors Geophysical parameters

Trappes Jan 2000–Dec 2003 Standard weather sensors 2 m aboveground

Temperature, humidity, and precipitation

SIRTA May 2004–Jun 2004 Standard weather sensors 15 m aboveground

Temperature, humidity, and precipitation

Radiometric station (BSRN standard) Downwelling direct, diffuse, and global solar irradianceRS90 and RS92 radiosondes Vertical profiles of temperature, pressure, humidity,

and horizontal wind; boundary layer heightBackscattering lidar Boundary layer height

FIG. 2. Root-mean-square difference between simulated tem-peratures and analyses (in the second model layer) in January1998. Large differences are observed at the center of the zoomedarea, showing that a certain degree of freedom is left to param-eterizations. Away from the zoomed area, differences are small,indicating that the large-scale component of the modeled circula-tion is adjusted toward ECMWF analyses.

APRIL 2007 C O I N D R E A U E T A L . 1477

Page 5: Assessment of Physical Parameterizations Using a Global Climate

al. 2005). For our study, we focus on an intensive ob-servation period associated with VAPIC that tookplace at SIRTA from 18 May to 17 June 2004. Duringthis period SIRTA measurements were enhanced bothin terms of sensors and operation hours with both dayand night measurements. Fifty Vaisala RS92 radio-sondes were launched from the SIRTA site in additionto the Vaisala RS90 radiosondes launched fromTrappes at the standard 0000 and 1200 UTC as part ofthe Meteo-France operational radiosonde network.Vertical profiles of thermodynamic properties mea-sured by radiosondes were used to derive the PBLheight. The SIRTA multichannel backscattering lidarfor cloud and aerosol research was operated on 20 dif-ferent days during VAPIC. The lidar data were ana-lyzed to retrieve the diurnal cycle of the PBL height onthose days. Direct, diffuse, and global components ofthe downwelling solar irradiance are measured atSIRTA according to protocols recommended by theBaseline Surface Radiation Network (Ohmura et al.1998). The two measurement sites are expected to berepresentative of a wide surrounding area, which isrelatively flat, with an homogeneous land surface di-vided in agricultural fields, housing and industrial de-velopments, and wooded areas. During the VAPIC in-tensive observation period, hourly observations of rela-tive humidity and temperature fields made by themeteorological center of Trappes are found to be con-sistent with SIRTA measurements. Thus, the spatialvariability of temperature and humidity is relativelyweak and its influence on 2000–03 monthly normals isdeemed negligible. Precipitation has a larger spatialvariability, and the average monthly data of five me-teorological stations, operated by Météo-France and lo-cated inside the grid cell of the model considered here,have been used to check the representativity of theTrappes station. The standard deviation of the 2000–03monthly normals for precipitation lies between 4.7%and 16.6%, which is less than the typical differencesbetween observed and modeled precipitation.

3. Sensitivity to soil scheme

The parameterization of land surface processes in cli-mate models is of great importance as it determinessensible and latent heat fluxes at the surface, which arethe lower boundary conditions for the energy and waterbudgets of the atmosphere. For the GCM, the key vari-able of land hydrology is evaporation because itstrongly controls partitioning of net incoming radiativeenergy into sensible and latent heat fluxes.

a. Numerical experiments

To evaluate the influence of the soil scheme, experi-ments with four different parameterizations are carriedout. In the reference experiment (ref), the soil schemeis SECHIBA, referred to above. In the simulation de-noted “b,” the SECHIBA scheme is replaced by asimple bucket model, representing the land surface as aone-layer soil reservoir, whose maximum water contentWmax is 150 mm. The soil moisture of the reservoir W iscomputed as the budget of precipitation P and evapo-ration E:

�W

�t� P � E, �2�

with all the water in excess of 150 mm being lostthrough runoff. Evaporation is computed as

E � �� |V |Cd qsat�Ts � � qs , �3�

where �, |V | , Cd, qsat, and qs are, respectively, the airdensity, wind speed, frictional drag coefficient, andsaturated relative humidity at surface air temperature,and relative humidity. Here � is an aridity function thatdepends on soil water content, as

��W� � min� W

Wmax �2, 1�.

Two additional simulations, “b5” and “b25” are runwith the same formulation for the aridity coefficient butwith a soil water content fixed to 5 and 25 mm, respec-tively (i.e., with � fixed to 1/15 and 1/3, respectively).

b. Mean seasonal cycles

The mean seasonal cycle of temperature and relativehumidity at 2 m are first compared with observations(see Fig. 3 for temperature and Fig. 4 for relative hu-midity). Annual cycles of air temperature and humiditysimulated by the model without setting � (ref and bsimulations) show more contrast than observed cycles.In particular, both parameterizations, SECHIBA andsimple bucket, simulate too high temperatures associ-ated with too low relative humidities during summerfrom June to September. It is noteworthy that this dryand warm bias in summer is a well-known deficiency ofthe LMDZ climate model over Europe (Hourdin et al.2006). Simulation b5 also exhibits a warm and dry biasduring summer. For b25, with artificially increasedevaporation during the warm period, temperatures arecolder and closer to observations during summer butthe relative humidity is overestimated.

Note that the b5 and b25 simulations were chosenhere for illustration of the sensitivity to surface evapo-ration. However, an intermediate value of � � 10/75

1478 M O N T H L Y W E A T H E R R E V I E W VOLUME 135

Page 6: Assessment of Physical Parameterizations Using a Global Climate

produces better results globally, as shown in the nextsection.

c. Water budget

The various simulations are associated with very dif-ferent evaporation rates (Fig. 5), a larger evaporationbeing generally associated with a colder and wetter at-mosphere. Those results also strongly suggest that theSECHIBA and bucket models both underestimate sur-face evaporation during summer.

As suggested by De Rosnay and Polcher (1998), thelow estimate of evaporation can be due to insufficientprecipitation in summer, inability of the soil to storeenough water during winter and/or to release waterduring summer, and excess of solar radiation. As shownin the last section, overestimation of solar radiation is abias of the LMDZ model and does not depend of thesoil scheme used. From consideration of the seasonal

cycle of soil moisture (Fig. 6) and evaporation (Fig. 5),it seems that the weak evaporation during summer withSECHIBA is explained by the soil being unable to re-lease water from a reservoir that is always above 70%of its maximum water-holding capacity. In comparison,the low level of soil moisture in the b simulation seemsto be the limiting factor for evaporation during sum-mer.

During winter and early spring, the bucket in the bsimulation is above half of its maximum content andthen evaporates at the potential evaporation rate (� �1). This high evaporation rate explains the cold and wetbias observed in spring and the subsequent rapid de-crease of the soil reservoir, leading to a very low evapo-ration in the following summer. It is the particular de-pendency of the � parameter that produces the phaselag between the ref and b simulations in terms of hu-midity and temperature (there is no such lag for the soilwater content).

The warm and dry summer bias in the ref, b, and b5

FIG. 4. Same as in Fig. 3, but for 2-m relative humidity. FIG. 6. Same as in Fig. 5, but for soil moisture.

FIG. 3. Comparison between observed (obs) and simulated (ref,b, b5, and b25, see text for details) seasonal cycle of 2-m tempera-ture.

FIG. 5. Simulated seasonal cycle of evaporation rate at a posi-tion corresponding to Trappes (48.8°N, 2.0°E). As described inthe text, simulations differ from each other with respect to the soilscheme employed.

APRIL 2007 C O I N D R E A U E T A L . 1479

Page 7: Assessment of Physical Parameterizations Using a Global Climate

simulations is also associated with an underestimationof precipitation rates (Fig. 7). Increasing the evapora-tion rate by setting the aridity coefficient to 1⁄3 (simu-lation b25) produces more precipitation, cancelling thebias in August and September but leading to an over-estimation of precipitation in June–July. During earlyspring, the larger evaporation rates in the b simulationalso produce more intense rainfalls. All simulationstend to overestimate rainfall in December and January.This overestimation is certainly partly due to a poorparameterization of large-scale precipitation processesin the model because local evaporation is expected tobe less important than advection at this season. Con-sidering vertical profiles in January (not shown here),the mean relative humidity simulated by LMDZ in thetroposphere is found to be higher than ECMWF analy-ses. This wet bias is certainly responsible, in addition toan imperfect representation of clouds, for the precipi-tation overestimate in wintertime. Moreover, LMDZinherits the defects of the ECMWF model as show inFig. 8.

It is clear in this example that we are testing morethan individual parameterizations. The larger springevaporation in the b simulation produces a larger rain-fall, which in turns increases the soil water content andsurface evaporation. This loop, which is at work in thefull climate GCM, can be assessed here with compari-son to in situ observations.

d. Diurnal cycles

The diurnal cycle amplitude is influenced by severalfactors, such as cloud cover and soil moisture (Stoneand Weaver 2003). By decreasing the downward solarradiation during the day and emitting longwave radia-tion during the night, clouds decrease the amplitude of

the diurnal cycle. Soil moisture is also expected to affectthe diurnal cycle by increasing both daytime surfaceevaporative cooling and nighttime greenhouse effectwarming. The correlation between the diurnal cycleamplitude and the parameterization of the land surfacescheme is examined in this section. Observed and simu-lated seasonal variations of the diurnal cycle amplitudeare displayed in Figs. 9 (temperature) and 10 (relativehumidity). The diurnal cycle amplitude is calculated asthe root-mean-square deviation of the hourly valuesfrom the daily average. Comparisons of the differentsimulations show that the diurnal cycle amplitude issensitive to the land surface parameterization and isstrongly anticorrelated with the evaporation rate (seeFig. 5).

The amplitude of the diurnal temperature cycle israther well simulated with the SECHIBA scheme.

FIG. 9. Same as in Fig. 3, but for the diurnal cycle of 2-mtemperature.

FIG. 7. Seasonal cycle of observed precipitation rates (graybars) and differences between simulated and observed precipita-tion rates (see text for details on the simulations).

FIG. 8. Seasonal cycle (years 2000–01) of observed precipitationrates (gray bars) and differences between simulated (by theECMWF model and the ref simulation) and observed precipita-tion rates.

1480 M O N T H L Y W E A T H E R R E V I E W VOLUME 135

Page 8: Assessment of Physical Parameterizations Using a Global Climate

There is however a phase lag associated with an under-estimation of the diurnal cycle in February–April andan overestimation in September–November. The bsimulation is satisfactory from June to January but di-urnal temperature amplitudes are too weak from Feb-ruary to May (months during which simulated meantemperatures are also too low). A consequence of usinga constant aridity coefficient is to decrease the annualvariation in diurnal temperature amplitudes. Differ-ences between b, b5, and b25 simulations confirm thatevaporation is a controlling factor of the diurnal tem-perature amplitude. A larger evaporation (b25) reducesboth the diurnal mean and the diurnal amplitude of thetemperature, leading to a large underestimation of thethermal cycle amplitude in summertime.

No model succeeds in catching the seasonal cycle ofthe diurnal amplitude for relative humidity. The aver-age amplitude is about right for ref, b5, and b25 simu-lations but with almost no seasonal cycle whereas ob-servations show an amplitude more than twice as largein summer than in winter. The simple bucket schemegenerates a seasonal variability but not in phase withthe observed one and with too small amplitudes in win-ter and spring.

4. Sensitivity to boundary layer scheme

In this section, we test, with the same model configu-ration, a new parameterization of the vertical transportin the convective PBL, based on a mass flux represen-tation of thermal cells.

a. Parameterization of the convective boundary layer

The convective boundary layer is characterized by aflux up the gradient (upward in a mixed layer which is

either neutral or more often slightly stable). This ver-tical transport is dominated by strong and organizedmesoscale motions (at the scale of the PBL itself).Those considerations have led in the past to a numberof adaptations or alternatives to the diffusive approach.Deardorff (1972) proposed introducing a countergradi-ent term to obtain upward heat fluxes even in a slightlystable atmosphere. This countergradient approach hasbeen refined further by Troen and Mahrt (1986) andHoltslag and Boville (1993). Stull (1984) proposed tocut radically with the diffusive approach by allowingdirect exchanges between all the layers in the PBL. Thetransilient matrix framework he proposed has led toseveral interesting attempts in terms of parameteriza-tions (Pleim and Chang 1992; Alapaty et al. 1997).

Mass flux ideas, originally developed for cumulusconvection, have led to various parameterizations ofthe PBL. The concept of mass flux approaches consistsof explicitly describing subgrid-scale vertical massfluxes used afterward to transport atmospheric quanti-ties such as potential temperature, water, momentum,or atmospheric constituents. For the convective PBL,the column is generally divided into an ascending buoy-ant plume, with mass flux f � �w, where , �, and ware the fractional cover, air density, and vertical veloc-ity in the thermal plume, respectively, and a compen-sating subsidence in the environment of mass flux �f.With these notations, the subgrid-scale turbulent flux ofa quantity � reads as

�w��� � f�� � ��, �4�

�1

1 � �f�� � ��, �5�

where � is the value of � inside the buoyant plume and� in its environment with � � � � (1 � )�. Theseequations can be considered as a general framework forthe mass flux application to the PBL.

This framework has been applied in the past essen-tially in the form of bulk models in which it is assumedthat, in the mixed layer, the various quantities are con-stant (Betts 1973; Lilly 1968; Randall et al. 1992) or varylinearly (Albrecht 1979; Wang and Albrecht 1990) as afunction of height. In those approaches, the mass fluxparameterization was used to prescribe an entrainmentflux at the top of the PBL. Bulk mass flux models weregenerally derived to analyze the physics of the PBL butsome were also used as real parameterizations for cir-culation models (Suarez et al. 1983).

The mass flux approach retained by Hourdin et al.(2002) differs from those previous studies by many as-pects. First, the thermal plume model does not intendto account for all the turbulent motions in the PBL. A

FIG. 10. Same as in Fig. 3, but for the diurnal cycle of 2-mrelative humidity.

APRIL 2007 C O I N D R E A U E T A L . 1481

Page 9: Assessment of Physical Parameterizations Using a Global Climate

diffusive formulation is kept in the model, which ac-counts in particular for small-scale turbulent transfersin the surface layer. Second, the mass flux is computedexplicitly from the buoyancy of warmer air taken in thesurface layer and transported through atmospheric lay-ers, to the level of neutral buoyancy. Overshoot is alsoexplicitly computed above that level. In that respect,the model is closer to some of the mass flux approachesretained for the representation of cumulus convection(Arakawa and Schubert 1974; Emanuel 1991; Tiedtke1989).

A similar two-stream framework in which a mass fluxcomputation coexists with a turbulent diffusion was al-ready proposed by Chatfield and Brost (1987). In thiswork, the mass flux profiles were prescribed as a func-tion of z/zi (where zi is the inversion height) based onsimilarity ideas. This parameterization was apparentlynot tested in circulation models. A parameterizationstill closer to the thermal plume model of Hourdin et al.(2002) has since been developed independently byA. P. Siebesma and J. Teixeira (2005, personal commu-nication) and tested by Soares et al. (2004).

Last, note that the idea of mass fluxes was applied byRandall et al. (1992) and Lappen and Randall (2001)who tried to combine the kinetic energy equation witha mass flux approach in order to derive closure rela-tionships for the mass flux and width of the thermals.The mass flux approach was also used by Abdella andMcFarlane (1997) to derive a third-order closure of theturbulent equations.

b. Numerical experiments

To evaluate the influence of the boundary layerscheme, we ran experiments with two versions of thePBL scheme. Both parameterizations were tested withthe SECHIBA surface scheme and the bucket modelwith a constant aridity coefficient � � 10/75. For bothcases, the vertical boundary layer flux of a quantity �can be written formally as

�w��� � ��K����

�z� �� � f�� � ��. �6�

Note that we use the classical approximation K 1 forthe mass flux parameterization.

In ref (same as in the previous section) and b10 simu-lations, the standard parameterization is used: the co-efficient K� depends on wind shear and the Richardsonnumber according to Laval et al. (1981), a constantcountergradient term �� � 1 K km�1 is applied forpotential temperature and the mass flux f is set to 0.

In the “ref � th” and “b10 � th” simulations, K� isdefined by parameterization 2.5 in Mellor and Yamada

(1974) based on a prognostic equation for the turbulentkinetic energy, the thermal plume model ( f � 0) and nocountergradient (� � 0). The mass flux f is computedwith a modified version of the thermal plume model ofHourdin et al. (2002) as described in the appendix.When compared with academic LES of the convectivePBL (Hourdin et al. 2002), a significant improvement isalready observed when using the Mellor–Yamadascheme alone (i.e., without the thermal plume model)instead of the standard LMDZ parameterization. Ver-tical profiles of potential temperature and heat flux arein good agreement with results of LES published byAyotte et al. (1996). In contrast, the standard param-eterization leads to unstable atmospheres when surfaceheating is too significant (the constant counter gradientterm being unefficient). The Mellor–Yamada schemealone nevertheless underestimates entrainment andleads to a too sharp inversion. Combining this diffusivescheme with the thermal plume model improves thevertical structure of the mixed-layer and top entrain-ment. An examination of the simulated thermals struc-ture indicates that heat is first supplied from the surfaceto the surface layer by the diffusion scheme and thentransported in the mixed layer by thermals.

c. Mean seasonal cycles

The use of a different parameterization of the bound-ary layer scheme does not produce significant differ-ences in mean temperatures (Fig. 11) and humidities(Fig. 12). The sensitivity to the PBL scheme is larger inwinter where the new parameterization produces aslightly wetter and colder climate. In wintertime, thethermal plume model is generally not active. Differ-ences between the standard and the new parameteriza-

FIG. 11. Comparison between observed (obs) and simulated(ref, ref � th, b10, and b10 � th, see text for details) seasonal cycleof 2-m temperature.

1482 M O N T H L Y W E A T H E R R E V I E W VOLUME 135

Page 10: Assessment of Physical Parameterizations Using a Global Climate

tion are consequently mainly due to the change in thediffusion scheme. An examination of vertical profiles inwinter indicates that local mixing in the surface layer ismore important with the Mellor–Yamada scheme andextends up lower in the troposphere. In summertime,mean temperature and relative humidity are less sensi-tive to the boundary layer parameterization than to therepresentation of land surface processes. The bestagreement with observations is thus obtained with thebucket model with a constant aridity coefficient � �10/75. The b10 � th simulation is somewhat better thanb10. The only difference in the water budget betweenthe two different boundary layer parameterizations(not shown here) is a slight increase in evaporationduring summer with the thermal plume model. Thisleads to an increased rainfall amount and higher soilmoisture.

d. Diurnal cycles

Figure 13 shows the differences of the mean seasonalcycle of the thermal diurnal cycle between model runsand observations. The annual cycle is strongly influ-enced by the PBL parameterization. Use of the massflux representation of thermal, associated with the Mel-lor–Yamada scheme for the turbulence exchange leadsto lower diurnal cycles from October to February andto higher cycles from March to September. TheSECHIBA land surface scheme results in overesti-mated diurnal cycles in spring and summer. The simplebucket model leads to a good agreement with observa-tions, in spite of an underestimation in July–August.Figure 14 displays the mean seasonal cycle of the diur-nal cycle of relative humidity. The new parameteriza-tion enhances the annual cycle with amplitude of thediurnal cycle higher in spring and summer. Comparedwith the b10 simulation, results of b10 � th are signifi-cantly improved with a more realistic annual cycle, but

the diurnal cycle is still underestimated in summer. Thestudy of hourly simulated temperature and relative hu-midity during the VAPIC observation period showsthat the improvement is due to both the Mellor–Yamada scheme and the thermal plume model. On theone hand, using the Mellor–Yamada parameterizationas the diffusion scheme produces a wetter nocturnalsurface layer. On the other hand, the thermal fluxmodel, by transporting humidity from the surface to themixed layer leads to a drier surface layer during day-time. Combining the Mellor–Yamada scheme and thethermal flux model produces a surface layer that is wet-ter during nighttime, drier during daytime, and inagreement with observations. The resulting 2-m rela-tive humidity cycle is consequently well caught. Theb10 � th parameterization is the most appropriate tosimulate the mean meteorological parameters (tem-perature and relative humidity) and the diurnal cycleamplitude. The warm bias of the reference experimentis significantly reduced, even if the model remains too

FIG. 14. Same as in Fig. 13, but for 2-m relative humidity.

FIG. 12. Same as in Fig. 11, but for 2-m relative humidity. FIG. 13. Same as in Fig. 11, but for seasonal variation of thediurnal cycle of 2-m temperature.

APRIL 2007 C O I N D R E A U E T A L . 1483

Page 11: Assessment of Physical Parameterizations Using a Global Climate

warm from April to July. The annual cycle of the diur-nal amplitude of temperature is in phase with the ob-served one and the simulated values are close to mea-surements.

5. Comparison with SIRTA observations

To illustrate the performances of the model with theoptimum parameterization (b10 � th), a 1-month pe-riod, corresponding to the VAPIC intensive observa-tion period, has been more precisely studied. Duringthis period, running from 18 May 2004 to 17 June 2004,observations performed by remote sensing and in situinstruments at the SIRTA observatory (weather sta-tion, cloud aerosol lidar, radiosonde measurements,and flux meters) are used to analyze simulated surfacetemperature, humidity, PBL height, and shortwavefluxes. We focus in this section on the b10 and b10 � thsimulations. The physical parameterization is the sameas previously. However, we decreased the time step forcomputation of meteorology (i.e., 1 min instead of 1.5)and physical processes (3 min instead of 30) to avoidnumerical instabilities in the computation of PBLheights.

a. Temperature and humidity

As shown in Fig. 15, simulated temperatures andrelative humidities are in satisfactory agreement with

observations. Amplitude of the diurnal cycle of humid-ity is enhanced with the thermal plume model, whereasthere are no significant differences in the two simulatedtemperatures. Daytime temperatures are sometimesoverestimated (21 and 26 May and from 30 May to 4June) and these periods are also associated with under-estimated relative humidities. Daytime humidities arenevertheless well captured when using the thermalplume model. Nighttime minimum temperatures areclose to observed data, whereas maximum humiditiesare generally underestimated by the model. In spite ofthe slight underestimation of the amplitude of the di-urnal cycle of humidity, the phase characteristics of thediurnal cycle are well captured by the model.

b. Boundary layer height

1) BOUNDARY LAYER HEIGHT DETECTION

Two methods are used to retrieve the PBL height.The first one is based on a dual-wavelength (532 and1064 nm) backscattering lidar and the other one makesuse of radiosonde profiles.

Aerosols in the atmosphere are efficient scatterers oflidar pulses. The sharp decrease of aerosol concentra-tion often observed at the interface between the PBLand the free troposphere results in a marked decreaseof the lidar backscattered signal at the top of the PBL.At night however, the PBL can be quite shallow while

FIG. 15. Observed and 2-m simulated temperature (T ) and relative humidity (RH) for (a) 18 May 2004–1 Jun2004 and (b) 2–17 Jun 2004. Measurements are performed 15 m above ground.

1484 M O N T H L Y W E A T H E R R E V I E W VOLUME 135

Page 12: Assessment of Physical Parameterizations Using a Global Climate

aerosols remain aloft in stratified layer called residuallayer. In this case, the sharp transition in the lidar back-scattered signal occurs at the top of the residual layer.In other situations clouds can reside at or near the topof the PBL, which causes the lidar signal to increaserather than decrease in the transition region. Morille etal. (2007) developed a multiple-test algorithm to iden-tify cloud and aerosol layers, to detect vertical zonesthat are nearly particle free, and to estimate the altitudeof the PBL top. This algorithm makes use of wavelettransform to identify singularities in the lidar backscat-tered profiles and searches coherence in the multipletests to identify the different particle layers. While theMorille et al. (2007) algorithm is efficient at estimatingthe PBL height in the presence of clouds, it is not ableto distinguish between the top of the boundary layerand the residual layer.

The method used to extract PBL height from radio-sonde measurements is a threshold method applied ona bulk Richardson number Rib(z) calculated as

Rib�z� �g�z � z0�

�z�

�z� � �z0�

u�z�2 � ��z�2 , �7�

where � is the potential temperature, g is the accelera-tion due to gravity, z is the height, z0 is the altitudereference (considered here as the first vertical pointavailable on the sounding profile), and u and � are thezonal and meridional wind components, respectively.The PBL height is estimated with a threshold value of0.21 (Vogelezang and Holtslag 1996).

2) RESULTS AND DISCUSSION

The boundary layer height, extracted from radio-sonde measurements and lidar retrievals, is comparedwith output of the model run in Fig. 16. The PBL heightis at its minimum throughout the night, increases to itsdaytime maximum because of diurnal heating, and thendecreases at sunset. For most of the days, the observeddaytime PBL is comprised between the altitude of neu-tral buoyancy zmix and top of thermal plume zmax (seethe appendix for the definition of zmix and zmax). Themodel succeeds in simulating the daily variability of thePBL pulse in response to different meteorological con-ditions. For instance, during the 25–27 May period, thePBL extends up to 2000–2500 m and the model fore-casts the maximum PBL height to be about 2800 m. Incontrast, the PBL height reaches only 1300 m on 3 June,600 m lower than the maximum predicted by the model(1900 m). The height of the mixed layer forecast by themodel is nevertheless slightly overestimated on thisday: simulated temperatures are also too high, whichcan explain a too efficient mixing. The reverse situationoccurs on 9–10 June with a PBL height predicted by themodel lower than the observed one. During this period,simulated temperatures are however in good agree-ment with observations. In general, the growth rate forthe mixing layer is well predicted by the model (espe-cially during the 23–27 May and the 14–17 June peri-ods). At night, the simulated PBL (“blh” in Fig. 16) isin good agreement with observations extracted fromradiosonde measurements. Note that zmix and zmax

FIG. 16. Comparison between observed and simulated PBL height (a) from 18 May to 1 Jun 2004 and (b) from2 to 17 Jun 2004. Observations are derived from the backscattering lidar (gray points), radiosondes measurementsfrom the meteorological center of Météo-France (square dots), and from the SIRTA observatory (triangle dots),zmix (black line) is the height of null buoyancy, zmax (gray line) the top of the plume (see the appendix for details),and blh (dashed line) the boundary layer height calculated as described by Mathieu et al. (2004).

APRIL 2007 C O I N D R E A U E T A L . 1485

Page 13: Assessment of Physical Parameterizations Using a Global Climate

are null throughout the night because thermals are notactive during this period. The simulation sometimeoverestimates the height of the very stable nocturnalboundary layer.

The “th” boundary layer scheme associated with thesimple bucket scheme with a constant soil moisture setto 10 mm is thus able to simulate the height of the PBLquite well during the intensive observation period.Some episodes are nevertheless not captured by themodel (e.g., overestimation on 9–10 June).

c. Shortwave cloud forcing

Figure 17 displays the shortwave cloud radiative forc-ing. The forcing is calculated with respect to the clear-sky shortwave radiation computed by the model. Dailyvariability is well captured by the model, in spite of asystematic underestimation of the simulated cloud ra-diative forcing. This is particularly true on days duringwhich temperatures are overestimated (21, 26, 31 Mayand 4 June). A poor estimation of the cloud fraction onthese days could explain the overestimation of theshortwave fluxes and, consequently, of the tempera-tures. Nevertheless, the cloud cover is improved whenusing the thermal plume model. This new parameter-ization significantly improves top entrainment (Hour-din et al. 2002) and vertical profiles of relative humid-ity. After having computed vertical transport with thethermal plume model, occurrence of oversaturation ischecked. In this case, condensation takes place and theLMDZ cloud scheme is used to diagnose the occur-rence of clouds. With this representation of dry convec-tion, more low and middle clouds are simulated, leadingto an increased cloud forcing. On 23 May and 5 and 16

June, no cloud cover is predicted by the b10 simulation,whereas the thermal plume model diagnoses a nonzerocloud forcing, close to the observations. But somecloudy episodes remain uncaptured with this param-eterization (e.g. 26 May).

6. Conclusions

In this study, we performed simulations using a vari-able resolution GCM, LMDZ, with different param-eterizations of land surface and boundary layer pro-cesses. The model was run with its zooming capabilityto reach 120-km resolution in the region of interest (thevicinity of Paris). One objective of this work was toadjust the parameterization schemes to produce thebest possible simulation (with no updates of initial con-ditions during the simulation period). Different landsurface schemes were first tested. The best agreementwith observations is obtained with the bucket modelwith a constant aridity coefficient � � 10/75. The com-putation of soil moisture as the budget of rainfall andevaporation (simulations “ref” and “b”) seems to havea seasonal inertia that is too weak leading to overlycontrasted 2-m temperature and humidity annualcycles. A second part was devoted to the sensitivityanalysis of the boundary layer parameterization. Thesimulated diurnal cycles are enhanced by using thethermal plume model associated with the Mellor–Yamada scheme for small-scale turbulence, providingbetter agreement with observations. This parameteriza-tion of the boundary layer leads to realistic seasonalcycles of the mean temperature and humidity as well asof their diurnal variations. The evaluation was first per-

FIG. 17. Observed and simulated shortwave cloud radiative forcing (CRFsw) (a) from 18 May to 1 Jun 2004and (b) from 2 to 17 Jun 2004.

1486 M O N T H L Y W E A T H E R R E V I E W VOLUME 135

Page 14: Assessment of Physical Parameterizations Using a Global Climate

formed on a 4-yr period (from 1 January 2000 to 31December 2003) and then on 1-month period (from 18May 2004 to 17 June 2004), corresponding to an inten-sive observation period. During this period, we com-pared hourly observations of the SIRTA observatorywith predicted temperatures, relative humidities, bound-ary layer heights, and shortwave fluxes. The “b10 � th”succeeds in catching most day-to-day variations.

This simulation will be used further as a basis forimproving the cloud scheme of the LMDZ model andof its coupling with the new thermal plume model. Inthis new representation of thermals (under develop-ment), the latent heat release inside the plume whenwater condenses is accounted for. The version used inthis study, developed for dry convection (Hourdin et al.2002), already improves the representation of boundarylayer clouds, but still underestimates the total cloudcover, which could significantly contribute to the warmbias observed in summer over the Paris area. The mea-surement of turbulent fluxes available since spring 2005at SIRTA will be of great help to further constrain thesurface scheme.

Besides the particular results obtained with LMDZ,the study presented here underlines the potentiality ofusing a climate model with a nudging and zooming ca-pability to work on physical parameterizations. Thisframework allows a fully coupled set of parameteriza-tions to be assessed on a given synoptic situation with avery low human and numerical cost. It is noteworthythat both the warm bias and the underestimation of theshortwave cloud radiative forcing are also observedwhen the same model (LMDZ) is used in a climatemode. Comparing day-by-day observations with modeloutputs allows us to identify which of the parameter-izations is responsible for these biases and to easilyassess new developments.

Acknowledgments. The author acknowledge A. Da-bas and J.-P. Aubagnac from the Centre National deRecherche en Météorologie for providing the radio-sondes launched from SIRTA during the VAPIC cam-paign and the high-resolution radiosonde data. We ac-knowledge Météo-France for providing radiosondedata from the departmental station of Trappes. Thanksare extended to the SIRTA team for providing pro-cessed geophysical parameters for our study.

APPENDIX

The Thermal Plume Model

The thermal plume model described by Hourdin etal. (2002) has been modified in order to simplify calcu-lations, to reduce central processing unit (CPU) time,

and to improve robustness. The physical processes in-volved stay unchanged, only the computation of thethermal plume characteristics is different.

A thermal plume is made of buoyant air coming fromthe unstable surface layer. The first step consists in de-termining the vertical mass flux f � �w in the thermalbetween each layer. The continuity equation for thethermal plume is written as

�f

�z� e � d, �A1�

where e represents the entrainment rate at which air issupplied to the base of the thermal plume and d is thedetrainment rate from the thermal into the environ-ment. The computation of the thermal properties is firstdone for d � 0. It is this first step that is modified in thenew version.

In the original version of the thermal plume modeldescribed by Hourdin et al. (2002) a buoyant plume iscomputed starting from each unstable layer near thesurface (a layer the potential temperature of which ishigher than that of the layer just above). Those inde-pendent plumes are then combined to produce a meanthermal that is used afterward to compute the verticaltransport of potential temperature, momentum, water,and tracers.

In the new version we directly compute the proper-ties of the mean thermal by imposing a priori the ver-tical distribution of the entrainment rate at the basis ofthe plume specified as e � �e* with ��

0 e* dz � 1. Theknowledge of the normalized entrainment e* is suffi-cient to determine the air properties and vertical veloc-ity inside the thermal plume. Parameter �, which is alsothe mass flux above the entrainment layer, is deter-mined afterward from a closure relationship.

The different steps of the calculation are the follow-ing.

1) Definition of the discrete entrainment rate inte-grated over the width �zk of layer k, E*k � e*k�zk, ineach unstable layer of the model (for which ���k ��k�1 � �k � 0). In the present simulations E*k ��max(���k, 0)�zk, where � is chosen in order tohave � E*k � 1.

2) Determination of the virtual potential temperature� and vertical velocity w inside the plume at eachlevel (� is the virtual potential temperature in theenvironment). Potential temperature is computedfrom the conservation equation:

l �

�k�1

k�l

E*kk

Fl�1�2c* , �A2�

APRIL 2007 C O I N D R E A U E T A L . 1487

Page 15: Assessment of Physical Parameterizations Using a Global Climate

where

Fl�1�2c* � �

k�1

k�l

E*k �A3�

is the conservative (computed with no detrainment)mass flux in the thermal. For the vertical velocityand for a stationary plume, the vertical variation ofthe kinetic energy must be equal to the work ofbuoyant forces:

12

wl�1�22 �

12 �Fl�1�2

c*

Fl�1�2c* wl�1�2�2

� gl � l

l�zl�1�2 � zl�1�2�. �A4�

The factor Fc*l�1/2 /Fc*

l�1/2 accounts for the fact that theair entrained in layer l enters the thermal plumewith a zero vertical velocity. It is then possible tocompute the height zmix of null buoyancy where thevelocity is maximal (wmax) and the top of the plumezmax, defined as the height where the vertical veloc-ity cancels. To reduce numerical instability zmax andzmix are defined as continuous and not discrete val-ues as was the case in the old version.

3) The closure consists in determining the coefficient�. From considerations of the geometry of thermalrolls in a 2D configuration, the entrainment rate isrelated to the velocity �k of lateral entrainment andthe width of one thermal cell L � rzmax (r being theaspect ratio of the cell) as

Ek � �E*k ��k zk�k

rzmax. �A5�

The closure relationship is based on the identity ofthe maximum vertical velocity in the thermal plumeand mean horizontal velocity in the entrainmentlayer:

wmax � �E*k�k. �A6�

Combining those two relationships finally leads to

� �wmax

rzmax �

Ek*2

�k zk

. �A7�

4) The true entrainment rate Ek � �E*k and conserva-tive mass flux Fc

k � �Fc*k can then be computed. The

detrainment and vertical transport are finally com-puted as in Hourdin et al. (2002).

REFERENCES

Abdella, K., and N. McFarlane, 1997: A new second-order turbu-lence closure scheme for the planetary boundary layers. J.Atmos. Sci., 54, 1850–1867.

Ackerman, T., and G. Stokes, 2003: The atmospheric radiationmeasurement program. Phys. Today, 56, 38–45.

Alapaty, K. A., J. E. Pleim, S. Raman, D. S. Niyogi, and D. W.Byun, 1997: Simulation of atmospheric boundary layer pro-cesses using local and nonlocal-closure schemes. J. Appl. Me-teor., 36, 214–233.

Albrecht, B. A., 1979: A model of the thermodynamic structure ofthe trade-wind boundary layer: Part II: Applications. J. At-mos. Sci., 36, 90–98.

Arakawa, R. A., and W. H. Schubert, 1974: Interaction of a cu-mulus cloud ensemble with the large scale environment, PartI. J. Atmos. Sci., 31, 674–701.

Ayotte, K. W., and Coauthors, 1996: An evaluation of neutral andconvective planetary boundary-layer parameterizations rela-tive to large eddy simulations. Bound.-Layer Meteor., 79,131–175.

Betts, A. K., 1973: Non-precipitating convection and its param-eterization. Quart. J. Roy. Meteor. Soc., 99, 178–196.

Bony, S., and K. A. Emanuel, 2001: A parameterization of thecloudiness associated with cumulus convection; evaluationusing TOGA COARE data. J. Atmos. Sci., 58, 3158–3183.

Chatfield, R. B., and R. A. Brost, 1987: A two-stream model ofthe vertical transport of trace species in the convectiveboundary layer. J. Geophys. Res., 92, 13 263–13 276.

Chiriaco, M., R. Vautard, H. Chepfer, M. Haeffelin, J. Dudhia, Y.Wanherdrick, Y. Morille, and A. Protat, 2006: The ability ofMM5 to simulate ice clouds: Systematic comparison betweensimulated and measured fluxes and lidar/radar profiles at theSIRTA atmospheric observatory. Mon. Wea. Rev., 134, 897–918.

Deardorff, J. W., 1972: Theoretical expression for the countergra-dient vertical heat flux. J. Geophys. Res., 77, 5900–5904.

De Rosnay, P., and J. Polcher, 1998: Modelling root water uptakein a complex land surface scheme coupled to a GCM. Hydrol.Earth Syst. Sci., 2, 239–255.

Ducharne, A., and K. Laval, 2000: Influence of the realistic de-scription of soil water-holding capacity on the global watercycle in a GCM. J. Climate, 13, 4393–4413.

Ducoudré, N., K. Laval, and A. Perrier, 1993: SECHIBA, a newset of parameterizations of the hydrologic exchanges at theland-atmosphere interface within the LMD atmospheric gen-eral circulation model. Climate Dyn., 6, 248–273.

Emanuel, K. A., 1991: A scheme for representing cumulus con-vection in large-scale models. J. Atmos. Sci., 48, 2313–2335.

Fouquart, Y., and B. Bonnel, 1980: Computations of solar heatingof the earth’s atmosphere: A new parametrization. Contrib.Atmos. Phys., 53, 35–62.

Gates, W. L., 1992: AMIP: The Atmospheric Model Intercom-parison Project. Bull. Amer. Meteor. Soc., 73, 1962–1970.

Ghan, S. J., L. R. Leung, and J. McCaa, 1999: A comparison ofthree different modeling strategies for evaluating cloud andradiation parameterizations. Mon. Wea. Rev., 127, 1967–1984.

Guichard, F., D. Parsons, J. Duhdia, and J. Bresch, 2003: Evalu-ating mesoscale model predictions of clouds and radiationwith SGP ARM data over a seasonal timescale. Mon. Wea.Rev., 131, 926–944.

——, and Coauthors, 2004: Modelling the diurnal cycle of deepprecipitating convection over land with cloud-resolving mod-

1488 M O N T H L Y W E A T H E R R E V I E W VOLUME 135

Page 16: Assessment of Physical Parameterizations Using a Global Climate

els and single-column models. Quart. J. Roy. Meteor. Soc.,130, 3139–3172.

Haeffelin, M., and Coauthors, 2005: SIRTA, a ground-based at-mospheric observatory for cloud and aerosol research. Ann.Geophys., 23, 253–275.

Hoke, J. E., and R. A. Anthes, 1976: The initialization of numeri-cal models by a dynamic-initialization technique. Mon. Wea.Rev., 104, 1551–1556.

Holtslag, A. A. M., and R. A. Boville, 1993: Local versus nonlocalboundary layer diffusion in a global climate model. J. Cli-mate, 6, 1825–1842.

Hourdin, F., P. Le Van, F. Forget, and O. Talagrand, 1993: Me-teorological variability and the annual surface pressure cycleon Mars. J. Atmos. Sci., 50, 3625–3640.

——, F. Couvreux, and L. Menut, 2002: Parameterization of thedry convective boundary layer based on a mass flux repre-sentation of thermals. J. Atmos. Sci., 59, 1105–1123.

——, and Coauthors, 2006: The LMDZ4 general circulationmodel: Climate performance and sensitivity to parametrizedphysics with emphasis on tropical convection. Climate Dyn.,27, 787–813.

Jeuken, A. B. M., P. C. Siegmund, L. C. Heijboer, J. Feichter, andL. Bengtsson, 1996: On the potential of assimilating meteo-rological analyses in a global climate model for the purposeof model validation. J. Geophys. Res., 101, 16 939–16 950.

Kasahara, A., 1977: Computational aspects of numerical modelsfor weather prediction and climate simulation. MethodsComput. Phys., 17, 1–66.

Krinner, G., and Coauthors, 2005: A dynamic global vegetationmodel for studies of the coupled atmosphere–biosphere sys-tem. Global Biogeochem. Cycles, 19, GB1015, doi: 10.1029/2003GB002199.

Lappen, C., and D. A. Randall, 2001: Toward a unified param-eterization of the boundary layer and moist convection. PartI: A new type of mass-flux model. J. Atmos. Sci., 58, 2021–2036.

Laval, K., R. Sadourny, and Y. Serafini, 1981: Land surface pro-cesses in a simplified general circulation model. Geophys.Astrophys. Fluid Dyn., 17, 129–150.

Lenderink, G., and Coauthors, 2004: The diurnal cycle of shallowcumulus clouds over land: A single-column model intercom-parison study. Quart. J. Roy. Meteor. Soc., 130, 3339–3364.

Lilly, D. K., 1968: Models of cloud-topped mixed layers under astrong inversion. Quart. J. Roy. Meteor. Soc., 94, 292–309.

Manabe, S., 1969: Climate and the ocean circulation. 1. The at-mospheric circulation and the hydrology of the Earth’s sur-face. Mon. Wea. Rev., 97, 739–774.

Marti, O., and Coauthors, 2005: The new IPSL climate systemmodel: IPSL-CM4. Tech. Note 26, IPSL, 82 pp. [Availableonline at http://dods.ipsl.jussieu.fr/omamce/IPSLCM4/DocIPSLCM4/.]

Mathieu, A., A. Lahellec, and A. Weill, 2004: Evaluation of anumerical weather forecast model using boundary layercloud top temperature retrieved from AVHRR. Mon. Wea.Rev., 132, 915–928.

Mellor, G. L., and T. Yamada, 1974: A hierarchy of turbulenceclosure models for planetary boundary layer. J. Atmos. Sci.,31, 1791–1806.

Morcrette, J. J., 2002: Assessment of the ECMWF model cloudi-ness and surface radiation fields at the ARM SGP site. Mon.Wea. Rev., 130, 257–277.

——, L. Smith, and Y. Fouquart, 1986: Pressure and temperaturedependence of the absorption in longwave radiation param-etrizations. Contrib. Atmos. Phys., 59, 455–469.

Morille, Y., M. Haeffelin, P. Drobinski, and J. Pelon, 2007:STRAT: An automated algorithm to retrieve the verticalstructure of the atmosphere from single-channel lidar data. J.Atmos. Oceanic Technol., in press.

Ohmura, A., and Coauthors, 1998: Baseline surface radiation net-work (BSRN/WCRP): New precision radiometry for climateresearch. Bull. Amer. Meteor. Soc., 79, 2115–2136.

Pleim, J. E., and J. S. Chang, 1992: A non-local closure model forvertical mixing in the convective boundary layer. Atmos. En-viron., 26A, 965–981.

Randall, D. A., Q. Shao, and C. Moeng, 1992: A second-orderbulk boundary layer model. J. Atmos. Sci., 49, 1903–1923.

Sadourny, R., and K. Laval, 1984: January and July performanceof the LMD general circulation model. New Perspectives inClimate Modelling, A. Berger and C. Nicolis, Eds., Elsevier,173–198.

Simmons, A. J., and D. M. Burridge, 1981: An energy and angu-lar-momentum conserving vertical finite-difference schemeand hybrid vertical coordinates. Mon. Wea. Rev., 109, 758–766.

Soares, P. M. M., P. M. A. Miranda, A. P. Siebesma, and J. Teix-eira, 2004: An eddy-diffusivity/mass-flux parametrization fordry and shallow cumulus convection. Quart. J. Roy. Meteor.Soc., 130, 3365–3383.

Stone, D. A., and A. J. Weaver, 2003: Factors contributing to di-urnal temperature range trends in twentieth and twenty-firstcentury simulations of the CCCma coupled model. ClimateDyn., 20, 435–445.

Stull, R. B., 1984: Transilient turbulence theory. Part I: The con-cept of eddy-mixing across finite distances. J. Atmos. Sci., 41,3351–3367.

Suarez, M. J., A. Arakawa, and D. A. Randall, 1983: The param-eterization of the planetary boundary layer in the UCLAgeneral circulation model: Formulation and results. Mon.Wea. Rev., 111, 2224–2243.

Tiedtke, M., 1989: A comprehensive mass flux scheme for cumu-lus parametrization in large-scale models. Mon. Wea. Rev.,117, 1779–1800.

Troen, I., and L. Mahrt, 1986: A simple model of the atmosphericboundary layer: Sensitivity to surface evaporation. Bound.-Layer Meteor., 37, 129–148.

Van Leer, B., 1977: Towards the ultimate conservative differencescheme: IV. A new approach to numerical convection. J.Comput. Phys., 23, 276–299.

Vogelezang, D. H. P., and A. A. M. Holtslag, 1996: Evaluationand model impacts of alternative boundary-layer height for-mulations. Bound.-Layer Meteor., 81, 245–269.

Wang, S., and B. A. Albrecht, 1990: A mean-gradient model ofthe dry convective boundary layer. J. Atmos. Sci., 47, 126–138.

APRIL 2007 C O I N D R E A U E T A L . 1489