Download as pdf or txt
Download as pdf or txt
You are on page 1of 29

Boundary-Layer Meteorology

https://doi.org/10.1007/s10546-020-00530-z

RESEARCH ARTICLE

Representation of Boundary-Layer Processes in Numerical


Weather Prediction and Climate Models

John M. Edwards1 · Anton C. M. Beljaars2 · Albert A. M. Holtslag3 ·


Adrian P. Lock1

Received: 26 November 2019 / Accepted: 15 May 2020


© Crown 2020

Abstract
Boundary-layer schemes are essential components of numerical weather-forecasting and
climate models. From simple beginnings 50 years ago, they have grown in sophistication
and detail. Here, we review development and discuss the key processes to be represented and
how they have most commonly been parametrized. We conclude by discussing the challenges
posed by ever-increasing model resolution and a growing emphasis on the forecasting of
extreme events. Throughout, we emphasize the place of the boundary-layer scheme within
the whole model and its interactions with other components of the model.

Keywords Planetary boundary layer · Parametrization · Numerical weather prediction ·


Climate modelling

1 Introduction

The modelling of the lower troposphere and its interaction with the underlying surface in
numerical weather prediction (NWP) and climate models is one of the most important appli-
cations of boundary-layer meteorology. Since the first numerical forecasts were made in the
middle of the twentieth century, the level of detail and sophistication of these models has
continually increased, resulting in a steady increase in the accuracy of forecasts (Bauer et al.
2015). These developments would have been impossible without the sustained growth in the
power and speed of digital computers that has occurred over the same period.
Numerical models are continuing to develop, with the atmosphere being represented at
ever-higher resolution. This increase in resolution raises a number of new modelling chal-
lenges and requires greater attention to the coupling between the planetary boundary layer
(PBL) and other components of the physical system. Societal factors, such as increased urban-
ization and the increasing importance of air-quality forecasting, are expanding the range of

B John M. Edwards
john.m.edwards@metoffice.gov.uk
1 Met Office, FitzRoy Road, Exeter EX1 3PB, UK
2 European Centre for Medium Range Weather Forecasts, Reading, UK
3 Meteorology and Air Quality Section, Wageningen University, Wageningen, The Netherlands

123
J. M. Edwards et al.

products required from models, as is the continuing development of Earth-system modelling


for studies of climate change.
An essential guiding principle in the development of numerical models has been fidelity
to the underlying physics of the atmosphere. Thus, in the present context, the requirements
of numerical forecasting have stimulated the development of boundary-layer meteorology,
while theoretical and experimental studies have made major contributions to the success
of numerical forecasting. Traditionally, the horizontal resolution of models has been much
coarser than the size of energy-containing eddies in the PBL, necessitating the introduction
of parametrizations to represent their net effects. There is no formal procedure for developing
parametrizations, rather the process depends on the synergistic combination of observational
campaigns, process studies and extensive analysis of behaviour in NWP and climate models
themselves, involving contributions by many scientists. An illustrative example is provided
by Randall et al. (2003), who describe how the process evolved in the context of cloud
modelling. The need for robust performance under computational constraints calls for a
degree of pragmatism. Not every proposed scheme has been implemented operationally,
even though its development may have yielded valuable insights.
Whilst the parametrization scheme adopted for the PBL is only one component of a numer-
ical model, it is of particular importance for the prediction of near-surface weather elements
that are of most direct relevance to the users of forecasts. In Fig. 1 we aim to convey a sense of
the capabilities of current models. Figure 1a shows the 10-m wind speed in a tropical cyclone
simulated in the current operational global configuration of the Met Office’s Unified Model.
The horizontal resolution (approximately 15 km ×10 km here) is such that the structure
of the eyewall is partially resolved, allowing more accurate predictions of the intensity of
storms than before, although the highest wind speeds are still often underestimated. There is
growing demand for forecasts of wind at higher levels in the PBL and Fig. 1b, reproduced
from Sandu et al. (2013), shows a comparison of the diurnal cycle of wind speed from the
Integrated Forecasting System of the European Centre for Medium-Range Weather Forecasts
(ECMWF) with observations from the tall 200-m tower at Cabauw in the Netherlands. The
model is able to simulate the diurnal variation of the wind speed quite well, although the
diurnal amplitude is somewhat underestimated at the higher levels. This is known to be due
to an overly diffusive and excessively deep stable boundary layer (SBL).
As well as the global configurations shown in the two preceding examples, forecasts are
also made at higher resolution on smaller domains, where the surface characteristics can be
represented in greater detail. Figure 1c and d show the 1.5-m temperature and visibility sim-
ulated using a limited-area configuration of the Unified Model in an area of complex terrain.
The 100-m resolution of this model is rather higher than that currently used operationally;
but it shows how the model is able to simulate the formation of cold pools in the valleys and
the associated development of fog.
The realistic simulation of boundary-layer depth is crucial for many purposes, such as
air-quality forecasting, and Fig. 1e shows the median value of the maximum boundary-layer
depth over the United States in summer, as simulated by the Integrated Forecasting System
in the ERA5 reanalysis (Copernicus Climate Change Service 2017). There is great variation
across the country, with a deeper PBL in the more arid western regions.
Finally, Fig. 1f, from a climate configuration of the Unified Model, shows a metric of
the local coupling between soil moisture and near-surface air temperature. It highlights
regions of strong coupling (blue), notably over North America and India, where local soil
moisture anomalies strongly influence the near-surface air temperature. Understanding such
couplings and ensuring that they are accurately represented is essential to the simulation of

123
Representation of Boundary-Layer Processes in Numerical…

(b)
(a)

(d)

(c)

(e)
(f)

Fig. 1 Examples of model output strongly dependent on the PBL scheme from the Integrated Forecasting
System and the Met Office’s system based on the Unified Model. a A tropical cyclone simulated in the current
operational global forecasting configuration of the Unified Model. Colours indicate the wind speed (m s−1 )
and contours the mean sea-level pressure. The scale, marked in intervals of 20 km, serves to indicate the
horizontal resolution of the model. b The annually-averaged diurnal cycle of wind speed at three levels in the
PBL, as simulated by the Integrated Forecasting System. Solid lines denote observations from Cabauw in the
Netherlands and dashed lines the corresponding simulated values. (Reproduced from Sandu et al. (2013) with
permission.) c The near-surface air temperature (K) simulated over an area of complex terrain using a research
configuration of the Unified Model. The scale, marked in intervals of 1 km, serves to indicate the horizontal
resolution of the model. Contours indicate the orography. d As c, but showing the near-surface visibility in
km. e The median value of the maximum height of the PBL (m) over the United States in summer 2019 (June
– August), as simulated in ERA5. f A metric of the local coupling between the near-surface air temperature
and the moisture content of the top 10 cm of the soil in summer (June–August) simulated using a climate
configuration of the Unified Model. The metric is calculated, following the LoCo methodology (Santanello
et al. 2018), as the product of the pointwise correlation coefficient between the daily average values of the two
fields and the standard deviation of the near-surface air temperature. Larger absolute values indicate stronger
coupling

123
J. M. Edwards et al.

climate change. We shall return to some of these topics in considering the future evolution
of boundary-layer parametrization.
In this review, we aim to trace the development of boundary-layer parametrization in
numerical models, showing how theoretical studies, observational campaigns, and appli-
cations have mutually reinforced each other. A brief historical sketch sets the context for
a discussion of how physical processes have been represented in typical schemes. Subse-
quently, we discuss some of the current modelling challenges and consider where progress
is needed to meet future modelling requirements. Necessarily, this review is not exhaustive
and further discussion may be found in the wide-ranging review of LeMone et al. (2019).

2 The Development of Boundary-Layer Schemes in Numerical Models

The first numerical weather forecasts (Bolin 1955) were made using the barotropic vorticity
equation to predict synoptic developments over the forthcoming few days. Early models had
a horizontal resolution of several hundred kilometres and treated the atmosphere as a single
layer. They did not account for diabatic processes, nor did they represent the PBL.
With the introduction of primitive equation models (Smagorinsky et al. 1965), it became
necessary to consider the lower surface and the PBL. The Kansas experiment of 1968 greatly
increased understanding of the lower levels of the PBL. Indeed, reviewing its legacy twenty
years later, Kaimal and Wyngaard (1990) concluded that it had provided “the first convincing
demonstration of the applicability of Monin–Obukhov similarity” theory (MOST) in the
surface layer. Above this, however, there was a “lack of knowledge about properties of the
PBL” (Deardorff 1972), and developing parametrizations required extrapolation from limited
knowledge.
At that time, computing resources were very limited, and two approaches to parametriza-
tion were suggested (Clarke 1970). One might attempt to resolve the PBL, albeit only very
coarsely, treating turbulence with a mixing-length model extrapolated from the surface layer.
Alternatively, recognising that the boundary layer might be shallower than a single model
layer, one might develop a bulk parametrization (Deardorff 1972). This approach had advan-
tages for the treatment of boundary layers capped by clouds and a sharp inversion, and was
successfully employed in the UCLA general circulation model (GCM) (Suarez et al. 1983).
However, it did require some method for predicting the boundary-layer depth, which was
not always well characterized (Louis 1979), and into the 1990s most models employed some
form of mixing-length scheme (Garratt 1993; Ayotte et al. 1996). A particularly influential
scheme was that introduced by Louis (1979), who combined an asymptotic mixing length
(Blackadar 1962) with a function of stability extended from the surface layer. Even today,
schemes of this type are still in operational use, especially to treat the stable SBL.
The more sophisticated and computationally expensive framework of second-order mod-
elling was introduced (Mellor and Yamada 1974; Yamada 1977; Mellor and Yamada 1982).
The simpler schemes of this type, involving a single prognostic of turbulent kinetic energy
(TKE), have often been used in research models, but operational implementations, such
as Benoit et al. (1989), have been rarer. Recently, however, improved TKE schemes, bene-
fiting from subsequent developments, have been adopted more widely, both in regional and
global models (Baldauf et al. 2011; Termonia et al. 2018; McTaggart-Cowan et al. 2019).
By the early 1980s, thanks to observational campaigns and modelling studies, the key
processes of the PBL were understood in outline, even though much remained to be done
in terms of quantification and gaining detailed understanding. Large-eddy simulation (LES)

123
Representation of Boundary-Layer Processes in Numerical…

was becoming viable as a tool for understanding the PBL and developing parametrizations
(Moeng 1984). Lilly’s introduction of the mixed-layer model (Lilly 1968) and the new per-
spective afforded by satellite imagery (Schubert 1976) clearly demonstrated the importance
of clouds in the PBL. The need for detailed in situ measurements to improve the modelling
of clouds in the PBL provided the motivation for major experiments involving aircraft, such
as The First ISCCP [International Satellite Cloud Climatology Project] Regional Experiment
(FIRE) (Albrecht et al. 1988). It was recognised that large coherent eddies, not amenable to
treatment by down-gradient diffusion, played an important role in the convective boundary
layer (CBL), and indeed could lead to counter-gradient transport (Deardorff 1966). More-
over, down-gradient diffusion, as in the Louis (1979) scheme, could not adequately represent
boundary-layer-top entrainment.
New ideas were proposed to represent large-scale transport. Bougeault and Lacarrère
(1989) introduced a mixing length into a TKE scheme based on parcel ascents [see also Cuxart
et al. (2000) and Lenderink and Holtslag (2004)]. Troen and Mahrt (1986) and Holtslag and
Boville (1993) proposed the alternative K-profile scheme, where a profile of diffusivity is
imposed across the whole PBL and an explicit non-gradient representation of large-scale
transport is introduced. Such schemes have proved very successful, convenient, and numeri-
cally robust for including more realistic representations of cloud processes and entrainment
(Lock et al. 2000; Hong et al. 2006) and were employed in many of the climate models partic-
ipating in the recent fifth phase of the coupled model intercomparison project, CMIP5 (Taylor
et al. 2012). Indeed, particularly in the context of climate modelling, realistic representation
of boundary-layer cloud and feedbacks is of crucial importance (Bony and Dufresne 2005).
More recently, parametrizations of coherent transport using the mass-flux approach that is
frequently applied to cumulus convection have been developed (Siebesma and Cuipers 1995;
Siebesma et al. 2007) and have been adopted by several modelling centres in the last few
years (Köhler et al. 2011; Han et al. 2016).
Over the past two decades, the need for more accurate and detailed forecasts has led to the
development of operational mesoscale models with a horizontal resolution finer than 10 km.
The traditional assumption that the model’s resolution is much coarser than the length scale
of eddies in the PBL no longer holds (Wyngaard 2004) and much current research is focused
on the implications of modelling in this so-called “grey zone.”
We turn now to the physical processes to be represented, but, before doing so, it is appro-
priate to emphasize the importance of experimental data for developing parametrizations.
Intensive periods of observation, as in the Cooperative Atmosphere Surface Exchange Study
(LeMone et al. 2000), facilitate detailed analysis of boundary-layer processes, but long-
term continuous monitoring of the PBL, such as that carried out at the Cabauw observatory
(Bosveld et al. 2020) and at the Southern Great Plains site of the Atmospheric Radiation
Measurement programme (Sisterson et al. 2016; Berg and Lamb 2016) is also invaluable.

3 Physical Processes and Their Representation

3.1 The Surface Layer

Boundary layers over land and ocean often have a fairly well-mixed structure, and the
leading-order approximation of the time evolution is controlled by the fluxes across the
top (entrainment) and at the surface. In a shallow near-surface layer, as in the capping inver-
sion, strong vertical gradients occur and the strength of mixing in these layers controls the

123
J. M. Edwards et al.

evolution of the mixed layer. It is therefore not a surprise that coupled and uncoupled atmo-
spheric models are very sensitive to the formulation of the transfer of momentum, heat and
moisture in the surface layer (e.g. Webster and Lukas 1992; Miller et al. 1992; Polichtchouk
and Shepherd 2016).
The surface layer is formally defined as the region where the height, z, is much smaller than
the depth of the PBL, h (Garratt 1994). In practice, the criterion is often taken as z/h < 0.1
(Stull 1988), although, as is often the case in asymptotic theory, solutions are useful for much
larger values of the parameter z/h. For instance, wind profiles at the Cabauw tower follow
similarity profiles well into the outer layer (although there are compensating effects; Holtslag
1984). This supports the pragmatic approach of many models in using rather thick surface
layers with bulk-transfer coefficients in the computation of surface fluxes,
Fφ = Cφ |Ul |(φl − φs ) , (1)
where Fφ is the kinematic flux of quantity φ (e.g., momentum, heat, moisture), Cφ is a
dimensionless transfer coefficient, |Ul | is wind speed at the lowest model level, φl is the value
of φ on that level, and φs is its value at the surface.1 Over the ocean, transfer coefficients are
of the order of 10−3 , but are substantially larger over land, dependent on land characteristics.
Early models used very simple empirical formulations for the transfer coefficients, with
different values for stable, unstable, land and ocean surfaces (e.g. Large and Pond 1982 for
ocean fluxes). Later, they were replaced by schemes based on MOST, where Cφ becomes
a function of zl /L (the height of the lowest level divided by the Obukhov length L) and
roughness lengths for momentum, heat and moisture transfer. The profiles used to derive
the transfer coefficients consist of a logarithmic part and a stability component (so-called
integral stability functions). Equation 1 has the disadvantage that it is an implicit formulation,
because Cφ is a function of zl /L that depends on fluxes (Stull 1988). However, the transfer
coefficients can easily be derived from the bulk Richardson number in an iterative way, as
in the Tropical Ocean Global Atmosphere (TOGA) Coupled Ocean–Atmosphere Response
Experiment (COARE) algorithm (Fairall et al. 2003), or by polynomial fitting, as proposed
by Li et al. (2010).
The MOST approach to parametrizing transfer coefficients is attractive for a number of
reasons: (i) MOST is well established for homogeneous surfaces and supported by many
observations, (ii) roughness lengths characterize the surface and are independent of stability,
provided that they are small compared to the Obukhov length, and (iii) in the constant-flux
layer, the description of the profiles below the lowest model level is numerically exact, and
the MOST profiles can be seen as finite elements. The latter is attractive because atmospheric
models have limited vertical resolution and the MOST formulation is robust even for poor
vertical resolution (Beljaars 1991). Furthermore, the profile functions can be used to inter-
polate between the surface and the lowest model level e.g. for post-processing of the 2-m
temperature. The integral profile functions are the integrated forms of the MOST gradient
functions, which have been observed in many circumstances. A comprehensive review is
given by Högström (1996). Different forms exist in the literature, reflecting the observa-
tional uncertainty, but the differences are fairly small compared to general parametrization
uncertainties in forecasting and climate models.
A popular set of functions is given in Beljaars and Holtslag (1991). These functions
have the characteristic that for every bulk Richardson number a value of zl /L can be found,
which is not the case for functions that imply a critical Richardson number. An important

1 In the analysis of field data, a transfer coefficient may more usefully be defined between two levels in the
atmosphere.

123
Representation of Boundary-Layer Processes in Numerical…

development has been the handling of the so-called free convection limit, when the large-scale
surface wind speed becomes very small. It was realized that large eddies in the boundary
layer maintain a non-zero near-surface wind speed that scales with the free convective velocity
scale w∗ = (h Tg w  θ  |0 )1/3 , where g is the acceleration due to gravity, T the temperature,
and w  θ  |0 the covariance of vertical velocity and potential temperature fluctuations at the
surface. So free convection is in fact a form of forced convection driven by large eddies
(Beljaars1995). This effect is often called gustiness and is included by replacing |Ul | in
Eq. 1 by |Ul |2 + (βw∗ )2 , where β is in the range 1–1.3. Including gustiness in the transfer
law has proved to be crucial, particularly in the warm-pool area of the Western Pacific, where
this scheme has a big impact on the general circulation (Miller et al. 1992; Polichtchouk and
Shepherd 2016).
Within the framework of MOST, research focused on the supporting parameters, namely
the roughness lengths. An important feature, which was not recognised in earlier models, is
that the roughness lengths for momentum, heat and moisture are not the same. The reason
is that form drag dominates momentum transfer, whereas scalars are ultimately transferred
by molecular diffusion. Over an aerodynamically rough surface, the aerodynamic roughness
length is typically about one tenth of the height of the roughness elements, while the rough-
ness lengths for scalars are generally one order of magnitude smaller; however, the thermal
roughness length in particular may be difficult to quantify (Mahrt and Vickers 2004).
Roughness lengths over the ocean are most often calculated using the formulation intro-
duced by Charnock (1955), whereby they are related to the surface stress, albeit with empirical
coefficients; the aerodynamic roughness length lies in the range 0.000025–0.01 m. At wind
speeds below 20 m s−1 there is a reasonable consensus, mainly through the large number of
ocean cruises in different projects (e.g. the TOGA COARE project, Webster and Lukas 1992).
They have resulted in a substantial dataset for optimization and refinement (see Brunke et al.
2003, for a comparison of algorithms). The TOGA COARE scheme is among the best. It
has evolved from one with a constant Charnock constant to one with a weak wind-speed
dependence to mimic wave effects, which may be represented directly if the atmospheric
model is coupled to a wave model (Janssen et al. 2002).
Fairall et al. (2003) conclude that the TOGA COARE scheme is accurate to within 5%
for wind speeds below 10 m s−1 and to within 10% scheme for wind speeds between 10
and 20 m s−1 . For higher wind speeds the situation is less clear, mainly because the direct
measurement of fluxes is virtually impossible. The topic is particularly important for the
simulation of tropical cyclones (see Sect. 4.2.6).
Over land, estimation of transfer coefficients is much more problematic than over the
ocean for two reasons: (i) land surfaces are often heterogeneous so MOST does not apply,
and (ii) the specification of roughness lengths is difficult, as it relies on global land-cover
datasets derived from satellite observations. Even for the only mildly heterogeneous site of
Cabauw in the Netherlands, Beljaars (1982) and Beljaars et al. (1983) showed that MOST
does not apply below 20 m. The very shallow layer near the surface, where the profiles are in
equilibrium with the local terrain, is only 0.1–0.2 m deep. However, above 20 m, where the
internal boundary layers of the heterogeneous terrain features have merged, the profiles are
logarithmic in the neutral limit and have a slope that corresponds with surface fluxes according
to MOST. To have a proper large-scale representation of the flow well away from the surface,
it is necessary to provide “effective roughness lengths,” such that the profiles extrapolated
from above 20 m match these roughness lengths. In fact, a roughness length is selected such
that the correct area-averaged surface flux is obtained, although the near-surface profile is not
necessarily realistic. This is fully consistent with the roughness-averaging concept proposed

123
J. M. Edwards et al.

by Mason (1988) and also supported by observations over hilly terrain (Grant and Mason
1990). The averaging of roughness over heterogeneous areas is very non-linear (Mahrt 1987),
and for momentum, the rough areas and obstacles such as tree rows dominate. For heat and
moisture the roughness lengths tend to be even smaller than open-terrain values because of
wind sheltering (Beljaars and Holtslag 1991).
Interpolation between the lowest model level and the surface is needed to calculate near-
surface temperature and humidity. For homogeneous terrain, MOST integral profiles are the
most suitable solution, but for heterogeneous terrain MOST is known to be less accurate.
Given the uncertainty of surface temperature over land and the poor terrain characterisation
in models, MOST or simplified versions are often used as a practical solution (Geleyn 1988).
The choice of roughness lengths in numerical models is still rather ad hoc. Tables have been
constructed to link roughness lengths to land-use maps (Dorman and Sellers 1989; Masson
et al. 2003), and they have been optimized on the basis of observations that are informative
of the related process e.g., near-surface wind speed (Sandu et al. 2011). Flow in canopies
(Finnigan 2000) and over hills (Wood and Mason 1993) have been studied intensively, but
the specification of parameters (e.g. effective roughness lengths) to characterize land sur-
faces on a global scale remains a formidable task that comes with large uncertainty. Simple
practical approaches, as described by Mason (1991) and Raupach and Finnigan (1988), are
still the solution of choice for the time being. More research is obviously needed to com-
bine detailed models for complex land surfaces, and satellite observations that provide the
necessary supporting information (e.g. Reigber and Moreira 2000). This is urgent not only
for forecast quality but also for data assimilation. Current systems, for example, have large
errors in radiative surface temperatures, because the land characteristics are not well defined,
and consequently window-channel skin temperatures cannot be assimilated (Balsamo et al.
2019).
Various papers question the usefulness of MOST for the parametrization of surface fluxes
(e.g. Mahrt and Vickers 2006; Mahrt 2008; Sun et al. 2016). Particularly for low wind speeds,
the vertical flux becomes intermittent and is dominated by mesoscale motions. Attempts have
been made to represent these effects using a mesoscale gustiness similar to that used in con-
vective situations (Mahrt and Sun 1995). However, no satisfactory complete parametrization
yet exists. The difficulty is that very little is known about the dependence of mesoscale gusti-
ness on the large-scale flow. This topic needs further research, e.g., by analyzing data from
LES over large areas.

3.2 Turbulence in the Outer Layer

Above the surface layer, the vertical kinematic turbulent flux, Fφ , can no longer be taken
as constant with z, and the depth of the PBL, h, becomes a relevant parameter: this region
is known as the outer layer. The simplest approach to parametrization here is to adopt a
first-order diffusive closure, writing,

∂ φ̄
Fφ = −K φ , (2)
∂z

where φ̄ is the resolved mean value of φ, and K φ is a diffusivity that can most easily be
specified using a diagnostic mixing-length model. In Prandtl’s original theory, developed for
pipe flows (Prandtl 1925), K φ = l 2 S, where l is the mixing length and S is the shear. In the
PBL, where the effect of atmospheric stability is crucial, it is usual to take l as the neutral

123
Representation of Boundary-Layer Processes in Numerical…

mixing length and to write the stability dependence explicitly as a function of the gradient
Richardson number, Ri; thus K φ = l 2 S f (Ri). This is known as the Louis (1979) scheme.
Near the surface, l = kz, where k is the von Karman constant. At greater heights, l becomes
limited, and this is most commonly represented by taking l as the harmonic mean of kz and one
or more additional length scales, following the approach of Blackadar (1962), there being no
formal theory to determine it. Often a single length scale, equal to a fraction of the boundary-
layer depth (Brown 1996), or even a specific value is chosen. The direct relation between
diffusivity and local shear in such schemes means that they are best suited to neutral or weakly
stable boundary layers, but less appropriate in the CBL, where turbulence is generated by
buoyancy. It is interesting to note that, for the fully turbulent SBL, the Louis (1979) scheme
has a strong basis in so-called “local scaling” as proposed by Nieuwstadt (1984) and revisited
by Derbyshire (1990) and Basu et al. (2006). This theory extends surface-layer similarity to
the outer layer by replacing surface fluxes by local fluxes in MOST.

3.2.1 Higher-Order Closure

An obvious approach to improving upon mixing-length models is to adopt a higher-order


turbulence closure. Mellor and Yamada (1974) considered the equations for the full second-
order moments of the turbulent velocity field and introduced a four-level hierarchy of closures
involving increasingly drastic approximation of its anisotropy. A crucial and somewhat uncer-
tain element of these schemes is the specification of a master length scale, for which Mellor
and Yamada (1974) adopted the neutral mixing length proposed by Blackadar (1962), but
included a length scale representing the vertical distribution of TKE. Subsequently, Yamada
(1977) suggested an approximation intermediate between the original levels 2 and 3, which
he termed “level 2.5:”2 in this case, a prognostic equation is retained only for the TKE, e,
and all other turbulent statistics are diagnosed. This has been the most extensively
√ explored
such closure. Here, the flux is still parametrized diffusively, but K φ ∝ e l K , where l K is
an appropriate length scale, proportional to the master scale, but incorporating the influence
of stability. Perhaps the most important feature of higher-order schemes is that they have the
potential of describing boundary-layer-top entrainment, by diffusing TKE from the bulk of
the boundary layer, where the production of TKE is concentrated, into the inversion where
it erodes the overlying stable layer.
Mellor and Yamada (1982) themselves remarked that the level-2.5 scheme was most
appropriate in stable and slightly unstable conditions, but was less consistent with their
assumptions in more unstable conditions. Indeed, it has often been found to yield a rather
shallow CBL (Hu et al. 2010; Xie et al. 2012) and this has led to the development of alternative
schemes for the CBL, as described in Sect. 3.2.2.
Recently, however, improved versions of the Mellor–Yamada scheme have been proposed.
Cheng et al. (2002) introduced improved formulations of the pressure–velocity and pressure–
temperature correlations, while Nakanishi and Niino (2009) both revised these formulations
and included a modified expression for the master length scale itself to account for atmo-
spheric stability. This improved scheme simulates more rapid growth of the CBL and is used
operationally by the Japan Meteorological Agency (Saito et al. 2007). A level-2.5 scheme
is used in the convective-scale configuration of the Consortium for Small Scale Modelling
(COSMO) model (Baldauf et al. 2011).
The emphasis in our review is not on cloud parametrization, but it is relevant to consider
whether turbulence schemes can support the representation of clouds by providing informa-

2 This approach is perhaps more commonly known as 1.5-order closure.

123
J. M. Edwards et al.

Fig. 2 Typical vertical profiles of (virtual) potential temperature, θv , and specific humidity, q, for a dry CBL
over land [modelled after Stull (1991) and adapted from Holtslag and Boville (1993) with permission]. The
arrows to the left indicate the specific humidity flux, Fq , and the arrows to the right, the heat flux Fθ . Also a
rising air parcel is indicated up to its intersection height, h c . The three regions are discussed in the text

tion about subgrid variability in temperature and moisture. This is quite difficult because
higher-order moments of turbulent quantities as well as variances are important, particularly
for cumuliform clouds (e.g. Bechtold et al. 1995). A major effort over the last 20 years has
been the development of the Cloud Layers Unified By Binormals (CLUBB) scheme, which
is based on higher-order turbulence closure and assumed probability density functions of
subgrid fluctuations of velocity, temperature and moisture (Golaz et al. 2002a, b; Larson and
Golaz 2005; Golaz et al. 2007; Larson et al. 2012). Depending on the selected complexity,
the turbulence model has seven to nine equations for second and third moments (Bogen-
schutz and Krueger 2013; Cheng and Xu 2006). Closure assumptions are made using double
Gaussian probability density functions. Higher-order closure has a long history in engineer-
ing flow (e.g. Lumley and Khajeh-Nouri 1975), but closure of the many complicated terms
(e.g. pressure–strain correlations) has always been problematic because of a lack of observa-
tions. With the increase of computer power and the maturing of LES models, simulations with
such models have now become an important source of pseudo-observations. The CLUBB
scheme has been successfully applied in various GCMs (e.g. Bogenschutz and Krueger 2013;
Guo et al. 2015). Guo et al. (2015) report that the transition from stratocumulus to cumulus
and from shallow to deep cumulus is simulated quite well, but that handling of mixed phase
clouds and ice microphysics remains a challenge.

3.2.2 Coherent Eddies and the K-profile Scheme

In the well-mixed CBL, schematically illustrated in Fig. 2, conserved quantities are trans-
ported mainly by coherent eddies spanning the entire PBL (e.g. Couvreux et al. 2010). Fluxes
depend on conditions throughout the CBL, not only on local mean gradients, which are small
in the middle of the CBL. Thus, whilst the flux, Fq , of specific humidity, q, is consistently
directed down the gradient of q, the kinematic heat flux, Fθ , is directed down the gradient of
potential temperature, θ , in regions 1 and 3, but not in region 2, because it is carried mainly
by parcels of air originating close to the surface that remain buoyant through most of the
depth of the CBL.

123
Representation of Boundary-Layer Processes in Numerical…

Fig. 3 Comparison of 54-, 60-, 66- and 72-h forecasts of 2-m specific humidity from two versions of the
ECMWF model with data from surface synoptic stations in the Mediterranean area. The Louis scheme is used
in Cycle 47 and the ECMWF version of the K-profile scheme in Cycle 48. It should be noted that a new
land-surface scheme was also introduced into Cycle 48. The impact in Cycle 48 compared to Cycle 47 is the
combined effect of higher moisture fluxes from the surface and entrainment at the boundary-layer top. The
averages are over all synoptic locations. Daily forecasts from 1200 UTC on 6 July to 1200 UTC on 22 July
1993 are used

Several approaches to this lack of locality have been proposed. Still within the framework
of a level-2.5 scheme, Bougeault and Lacarrère (1989) introduced a parametrization of l K
based on parcel ascents that reflects the overall structure of the CBL3 . Lenderink and Holtslag
(2004) and Rodier et al. (2017) have suggested further refinements to include the effect of
local stability and wind shear. This alone is insufficient to represent counter-gradient fluxes
in region 2, but a slightly unrealistic unstable stratification may be tolerable, or a gradient
adjustment may be applied (e.g., Bougeault and Lacarrère 1989).
A more radical approach is adopted in the K-profile scheme (Troen and Mahrt 1986; Holt-
slag and Boville 1993; Hong and Pan 1996; Lock et al. 2000; Hong et al. 2006). Recognizing
that the CBL is often capped by a pronounced inversion and has a well-defined depth, h,
an explicit non-gradient flux, FφNL = Fφ |0 f NL (z/h) is included in Eq. 2. Here, Fφ |0 is the
surface flux of φ and f NL is a prescribed function. Although there is no complete theory
to predict f NL from first principles, feasible parametrizations can be derived by applying
dimensional analysis to results from LES models (e.g. Holtslag and Moeng 1991). Generally,
h is diagnosed from a parcel ascent (Beljaars and Viterbo 1998; Lock et al. 2000) or using a
parcel ascent and a criterion on the bulk Richardson number (Troen and Mahrt 1986; Holt-
slag and Boville 1993; Vogelezang and Holtslag 1996). The local diffusivity, K φ , is usually
represented diagnostically as the product of a velocity scale for the whole CBL, h, and a
profile in z/h, again derived from large-eddy simulations, though it may also be determined
using a TKE-based closure. The most important non-gradient fluxes are those of scalars, but
the non-gradient transport of momentum is also significant (Frech and Mahrt 1995; Noh et al.
2003; Brown et al. 2008).

3 This may be described as a non-local mixing length, but the term “non-local” is often used to describe fluxes
not proportional to the local gradient. To avoid confusion, we will use the term “non-gradient” in the latter
sense.

123
J. M. Edwards et al.

Many studies have shown that such schemes yield a deeper, better-mixed and more realistic
CBL than purely local alternatives (Holtslag and Boville 1993; Beljaars and Viterbo 1998;
Hu et al. 2010; Xie et al. 2012).

3.2.3 K-profile Extension to Include Cloud-Top-Driven Mixing

Surface heat fluxes and wind shear are not the only processes that can generate turbulence in
the PBL. Of particular importance to both NWP and climate simulations is the stratocumulus-
capped PBL, via its huge local impact on radiative fluxes and thence the surface and global
energy budgets. Turbulence in stratocumulus is generated primarily by radiative cooling
at the cloud top (the additional role of cooling from evaporation of cloud droplets when
mixed with entrained dry air has been the subject of some controversy, but recent very high-
resolution simulations (Mellado 2017) suggest a more muted role in enhancing the radiative
generation). This leads to negatively buoyant thermals sinking through the PBL. By analogy
with turbulence driven by surface heat fluxes, several authors have introduced an additional
K-profile scaled primarily by cloud-top radiative cooling (Köhler et al. 2011).
There remains a significant challenge to diagnose the vertical extent of cloud-driven mix-
ing, otherwise permitting cloud-top mixing to penetrate to the surface could be self-sustaining
(Paluch et al. 1994). Lock et al. (2000) used a diagnostic parcel descent from the cloud top, but
this was superseded (Lock 2001) by a method that iteratively adjusted the vertical extent of
the K-profiles to ensure that, from the resultant buoyancy flux, the magnitude of the buoyancy
consumption of TKE was limited to a specified fraction of buoyancy production, integrated
across the boundary layer (Turton and Nicholls 1987; Stevens 2000). This allowed a realistic
decoupling of the cloud layer from the surface, in the sense that surface heat and moisture
fluxes are no longer efficiently mixed into the cloud layer, and so this typically initiates a
thinning of the cloud layer and ultimately a transition to trade cumulus (e.g., Bretherton and
Wyant 1997).

3.3 Entrainment

While the use of non-gradient flux profiles and the shape of the K-profile affect the vertical
structure of the PBL, the time evolution of the unstable PBL is largely governed by the fluxes
across the surface and across the PBL top. As a result, models are found to be very sensitive
to their representation of entrainment, in terms not only of how it is formulated, but also of
how well it is resolved on the vertical grid (Ayotte et al. 1996). This motivated several authors
to include an explicit parametrization for the entrainment flux, separate from the generation
of mixing within the PBL. This methodology has been applied in both higher-order closures
(e.g., Grenier and Bretherton 2001) and particularly in K-profile schemes (Beljaars and Betts
1992; Lock et al. 2000), sometimes combined with an additional non-gradient entrainment-
flux profile (Noh et al. 2003). However, there is a wealth of literature proposing alternative
parametrizations of the entrainment fluxes, based on both observational data and LES. Even
for the cloud-free PBL, where these are generally based on bulk TKE-budget arguments
(Tennekes and Driedonks 1981), there is significant disagreement, often arising from how to
incorporate the effects of wind shear both within and across the top of the PBL, and leading
to significant impacts on the PBL evolution (e.g. Conzemius and Fedorovich 2006).
For the cloud-capped PBL, entrainment parametrizations have been derived from aircraft
data (Nicholls and Turton 1986) and LES (Lock 1998; van Zanten et al. 1999; Moeng 2000;
Duynkerke et al. 2004). Typically these are also based on bulk energetic arguments, but the

123
Representation of Boundary-Layer Processes in Numerical…

interactions between entrainment and the cloud make for significant complications, such that
different parametrizations can lead to very significant differences in cloud evolution (Stevens
2002). Also in the TKE-based turbulence closure, it is non-trivial to find a suitable length scale
formulation in the inversion such that observed entrainment is reproduced (Lenderink et al.
1999). A very good review of these complications, such as the impact on the evaporative
cooling from interactions between turbulent entrainment and cloud microphysics and the
potential role played by cloud-top wind shear, is given in Mellado (2017).
Furthermore, there are important numerical difficulties in representing the sharp interface
between the turbulent boundary layer and the quiescent warm and dry air above. Lenderink
and Holtslag (2000) illustrated how this could lead to spurious “numerical” entrainment when
there is large-scale subsidence (as is typically the case for stratocumulus layers). Prompted
by this and similar issues with the grid-scale interaction between turbulent and radiative
fluxes across the inversion (Stevens et al. 1999), Grenier and Bretherton (2001) developed
the concept of reconstructing the subgrid inversion structure to allow a realistic coupling of
the turbulence with radiative fluxes and the resolved dynamics. This idea was also devel-
oped by Lock (2001), who showed the importance of the correction for subtropical marine
stratocumulus in a GCM (Lock 2004), but has subsequently received little attention.
The important role entrainment plays in the PBL is illustrated in Fig. 3, which shows the
significant reduction of the errors in forecasts of near-surface specific humidity that resulted
from the introduction of a parametrization of entrainment into the ECMWF model in 1993.

3.4 The Eddy-Diffusivity Mass-Flux Approach

Although the parametrizations of non-gradient fluxes described in Sect. 3.2.2 perform well
in dry convective or stratocumulus-capped boundary layers that are forced by fluxes from
clearly defined boundaries, it is not obvious how they should be extended to cases where con-
densational heating becomes a major source of buoyancy. Moreover, convective clouds grow
from coherent thermals in the CBL (LeMone and Pennell 1976) and transport in convective
clouds has widely been parametrized using the mass-flux concept. This suggests that it might
be advantageous to adopt this approach in the CBL and to write FφNL = (M/ρ)(φu − φ̄),
where M is the mass flux, ρ is the density, and φu is the average of φ across the coherent
updraught (Siebesma et al. 2007). A plume model is now required to provide equations for
M and φu ; extensions to include multiple plumes have been described by Sušelj et al. (2013).
This more unified eddy-diffusivity mass-flux approach provides a realistic representation
of the entrainment layer and ventilation at the top of the PBL when this is generated by
updraughts (Siebesma et al. 2007), and may be more appropriate for simulating the transition
from stratocumulus to cumulus-dominated boundary layers. A number of forecasting centres
have recently implemented such schemes (e.g. Köhler et al. 2011; Han et al. 2016; Termonia
et al. 2018; Olson et al. 2019). Similar issues concerning internal sources within the PBL
arise in air-quality modelling, where a related mass-flux parametrization for non-gradient
fluxes has been developed by Pleim (2007).

3.5 The Stable Boundary Layer

Turbulence in the SBL is generated by shear production, so it might be expected that the
SBL could be represented well by a simple local closure. In fact, parametrization of the SBL
has proved to be a rather intricate problem (Holtslag et al. 2013; Sandu et al. 2013) and
key forecast products, such as near-surface air temperature, are very sensitive to changes

123
J. M. Edwards et al.

in land-surface or atmospheric parametrizations, including that of the PBL. Whilst local


scaling (Nieuwstadt 1984) provides a strong theoretical basis for the parametrization of the
fully turbulent SBL, the very stable boundary layer with intermittent turbulence is less well
understood (Grachev et al. 2005). At high stability, non-turbulent processes assume greater
importance and the performance of PBL parametrizations must be evaluated in this broader
context.
In very stable conditions wave-like motions may occur and interactions between such
motions and turbulence have been observed (Cava et al. 2015). Moreover, measurements
of fluxes in weakly turbulent conditions are difficult and uncertain (Wilson et al. 2002),
while LES of the weakly turbulent SBL remains challenging (Chung and Matheou 2014).
Such problems have often frustrated the development and introduction of more sophisticated
parametrizations into operational models. Indeed, the risk that the surface might become
decoupled from the atmosphere (Louis 1979) has led to the operational use of parametrizations
with rather high levels of turbulent mixing (Cuxart et al. 2006), though it should be noted
that improvements in other areas of parametrization have allowed some reduction of this
enhanced mixing (e.g. Brown et al. 2008).
The overall importance of non-turbulent motions for large-scale modelling is not yet
known, but new theoretical approaches to their parametrization are being developed. Suko-
riansky et al. (2005) describe a spectral closure derived from first principles, quasi-normal
scale elimination (QNSE), that can be formulated as a K – turbulence model. The inclusion
of breaking internal gravity waves in QNSE prevents the eddy viscosity from vanishing at
very high stability. Quasi-normal scale elimination is available as a boundary-layer scheme
in the Weather Research and Forecasting (WRF) model (Tastula et al. 2014). Mauritsen et al.
(2007) describe a closure based on the total turbulent energy that includes the potential energy
due to density fluctuations in the fluid, instead of a more conventional closure based on TKE
alone. Buoyant destruction of TKE is then interpreted as a transfer to potential energy. Sim-
ilar to QNSE, this scheme also allows stresses to remain finite at high stabilities, while the
heat flux tends to zero. It has been adopted in ICON-A, the atmospheric component of the
icosohedral nonhydrostatic Earth-system model (Giorgetta et al. 2018).
Gravity waves generated by small-scale surface orography may also augment momentum
mixing and a parametrization has been suggested and assessed by Steeneveld et al. (2009)
and Tsiringakis et al. (2017). This parametrization has been adopted operationally for high-
resolution forecasting in the U.S.A. (Olson et al. 2019).
High vertical resolution appears important for the simulation extremely shallow and stable
boundary layers, such as those encountered over the Antarctic Plateau, particularly in winter
(Genthon et al. 2013). A current intercomparison, the fourth phase of the GEWEX [Global
Energy and Water Exchanges] atmospheric boundary layer study, GABLS4, focusing on the
slightly less extreme conditions of the Antarctic summer, demonstrates that such conditions
are challenging for large-eddy models and that high vertical resolution is required (Couvreux
et al. 2020). van der Linden et al. (2019) have shown that a vertical resolution as fine as
0.08 m is desirable for LES of the Antarctic winter SBL. Additionally, mesoscale modelling
studies (Vosper et al. 2013) suggest that fine vertical resolution near the surface is beneficial
for the simulation of processes such as cold-air pooling. Strictly, at high vertical resolution,
the roughness sublayer must be properly considered (see Sect. 4.2.3).

123
Representation of Boundary-Layer Processes in Numerical…

4 Current and Evolving Challenges

Thus far, we have discussed the traditional problems of boundary-layer parametrization.


While progress continues to be made in these areas, evolving forecasting requirements, such
as those arising from the adoption of ever-finer computational resolution and a growing
emphasis on extreme events, pose additional challenges; we now consider some of these
issues.

4.1 Higher Resolution: The Grey Zone and the Exascale Challenge

Throughout the history of NWP and climate modelling, the continual growth in computing
power has enabled physical processes to be represented with ever higher resolution and in
greater detail to meet the demand for more accurate and comprehensive forecasts. Today,
global forecasting models are run with horizontal resolutions of about 10 km (Neumann et al.
2019). At finer resolutions, deep convection begins to be resolved and it becomes feasible
not to parametrize this process. Whereas models with parametrized convection typically
exhibit a peak in convective precipitation around noon, convection-permitting kilometre-
scale models show a more realistic delayed peak in precipitation and better spatial statistics
of its distribution (Prein et al. 2013). On shorter forecast ranges and over restricted domains,
such models are run operationally with resolutions approaching 1 km (Baldauf et al. 2011;
Tang et al. 2013). Consequently, there is an aspiration to develop global forecasting models
with a resolution of about 1 km (Wedi 2014; Neumann et al. 2019). Roberts et al. (2018) have
demonstrated that atmospheric processes in climate models are better represented when the
resolution of those models is increased. Also, with increasing urbanization, the representation
of urban areas is becoming a priority (Boutle et al. 2016; Baklanov et al. 2018), motivating
the development of sub-kilometre city-scale models.
Whilst the grid size of forecasting models has hitherto been much coarser than the length
scales of the energy-containing eddies in the PBL, requiring boundary-layer processes to be
parametrized completely, at the resolutions now being used for regional NWP and envisaged
for the next generation of global models, these two scales become comparable. This situation
should be distinguished from that of LES, where the grid length is much smaller than the
scale of the energy-containing eddies, which are fully resolved.
The representation of the PBL in this intermediate “grey zone” raises a number of issues for
boundary-layer schemes (Wyngaard 2004) and is an active area of research. In the grey zone,
coherent energy-containing eddies in the PBL are partially resolved, but their interaction
with smaller-scale motions that still contain significant energy is not represented, so they
are not fully realistic surrogates for eddies, either in the real atmosphere or LES, and their
appropriateness as features of the simulation has been questioned (Ching et al. 2014). This
issue has not yet been resolved (Kealey et al. 2019).
Most parametrizations for the grey zone have been developed by seeking to partition
the turbulence between resolved and subgrid components, guided by LES (Honnert et al.
2011). If a conventional one-dimensional parametrization is applied in the grey zone, the
resolved motions are too weak and the one-dimensional scheme is too active, whilst if the
one-dimensional scheme is omitted entirely, the resolved motions are too active (Honnert
et al. 2011). The one-dimensional parametrization must therefore be retained, but the fluxes
it predicts should be scaled by a factor that depends on the ratio of the PBL depth to the
grid resolution (Boutle et al. 2014; Shin and Hong 2015; Honnert 2016), thereby allowing
the resolved motions to become stronger (such a scheme is described as scale-aware). In

123
J. M. Edwards et al.

Mellor–Yamada schemes, the master length scale is reduced in the grey zone (Ito et al.
2015). Moreover, the parametrization should encompass three-dimensional effects, such as
horizontal shear and transport.
Not only must the subgrid model become scale-aware in the grey zone, but the surface
boundary condition also becomes a serious issue. Monin–Obukov similarity theory applies
to averaged quantities, and does not provide a relation for the time-dependent surface flux
in relation to the forcing by large eddies. Stoll and Porté-Agel (2006) review and compare
different approaches with the goal of reproducing MOST profiles in the surface layer with
LES. A fundamental problem is that the small-scale eddies in the surface layer interact directly
with the large-scale outer eddies. In LES, Mason and Thomson (1992) address this issue with
a stochastic back-scatter scheme. In mesoscale models, the issue is yet to be resolved.
Delivering forecasts at such high resolution will require computers that operate in the exas-
cale range (1018 floating-point operations per second). It has been argued that fundamental
physical limitations on current computing hardware have been reached, restricting the poten-
tial for further increases in performance by conventional means, and that the forthcoming era
of exascale computing will be characterized by a wider variety of computer architecture and
increasing software complexity (Lawrence et al. 2018). These factors will pose challenges
to the process of model development.

4.2 Parametrizations Strongly Coupled to the PBL Scheme

There is an increasing focus on the forecasting of natural hazards and on understanding


how their frequency and severity may change in a changing climate (Sillmann et al. 2017).
Many natural hazards are compound events (Zscheischler et al. 2018), involving interactions
between different processes that are best represented in coupled models. Moreover, changes
in extreme events may be more pronounced than those in the mean climate (Donat et al.
2017). Together with the ever-increasing resolution of models, these considerations raise the
importance of understanding and improving the interaction between parametrizations of the
PBL and those of related processes.
The connection between the PBL and convection schemes has already been noted in
Sect. 3.4 and the connection with the cloud scheme is implicit in the discussion in Sect. 3.3,
but interactions with the surface are of especial importance. The understanding and quan-
tification of coupling between the land surface and the atmosphere (e.g. Fig. 1f) is a topic
that is currently attracting considerable interest from modellers (Santanello et al. 2018) and
experimentalists (Wulfmeyer et al. 2018). These interactions and the desire to represent the
surface in greater detail present complex challenges; here we can only summarize the issues.

4.2.1 Complex Terrain

Complex subgrid terrain is extremely hard to handle in large-scale models. Increasing model
resolution is the most obvious way forward, but heterogeneity exists at all scales and a
subgrid parametrization will be needed in many situations. Currently, the effective parameter
approach is rather popular, as it has some justification for fairly small horizontal scales and
strong winds (Beljaars 1982; Beljaars et al. 1983; Mason 1988). The specification of an
“effective roughness length” allows the simulation of realistic surface fluxes and leads to
realistic logarithmic profiles well above the heterogeneity. The concept also applies to hilly
terrain where, in strong-wind conditions, observed profiles well away from the surface have

123
Representation of Boundary-Layer Processes in Numerical…

a logarithmic shape, with a slope that corresponds to the surface stress including the form
drag of the hills (Grant and Mason 1990).
In mountainous terrain and weaker winds, the PBL may have a complex multi-layered
structure and is much affected by slope flows (Lehner and Rotach 2018). Field measurements
over steep slopes have shown that, not surprisingly, the profiles of wind speed and temperature
do not obey MOST (Nadeau et al. 2013). Traditional parametrizations developed for flat
terrain are therefore inadequate, but parametrizations appropriate to complex terrain are only
beginning to be developed.
Goger et al. (2018) evaluated the performance of the turbulence parametrization of the
COSMO NWP model with a horizontal grid spacing of 1.1 km for the complex Inn Valley
in Austria. Using an improved turbulence parametrization including horizontal exchange
processes, the modelled TKE has a more realistic structure in comparison with field data.
This is especially true in situations where the TKE production is dominated by shear related
to the afternoon up-valley flow, and during nights, when an SBL is present.
At high resolution, the terrain-following coordinate systems typically used in operational
models become less accurate and alternative numerical techniques, such as the immersed
boundary method (Lundquist et al. 2012), are being developed.
The existence of great variability on small scales in such a PBL raises challenges in
validating such schemes, and novel observational approaches (e.g. Cuxart et al. 2019) will
surely contribute to future progress.

4.2.2 Fog

The forecasting of fog is a notoriously difficult problem, involving the interaction between
weak turbulence, cloud microphysics, radiation and the underlying surface, and is often
influenced by drainage flows (e.g. Fig. 1c, d) and dew deposition (Price et al. 2018). Bergot
et al. (2007) showed that the forecast life cycle and characteristics of radiation fog are
very sensitive to the detailed formulation of the model. Steeneveld et al. (2015) evaluated
the performance of two operational mesoscale models (namely HARMONIE and the WRF
mesoscale models) for two contrasting warm fog episodes over the relatively flat terrain
around the Cabauw tower facility in the Netherlands. It appears that the boundary-layer
formulation is critical for the forecasting of fog onset, while for fog dispersal, the choice of
the microphysical scheme is a key element.

4.2.3 The Roughness Sublayer and Plant Canopies

Historically, the lowest levels of atmospheric models have been well above the roughness
elements of the surface, such as plant canopies, and the effect of the surface could be repre-
sented simply by using a roughness length. With increasing vertical resolution, models are
now beginning to resolve the roughness sublayer, where the flow is affected by the structure
of the canopy and deviates from MOST (Finnigan 2000). Theoretical understanding empha-
sizes the importance of the shear layer at the top of the canopy. Parametrizations consistent
with this theory are beginning to be developed (Harman and Finnigan 2007, 2008).

4.2.4 Urban Areas

Urban areas present a number of distinct modelling problems, including sharp changes in
roughness, multiple local internal boundary layers, a deep roughness sublayer and anthro-
pogenic heat sources (Martilli 2007; Barlow 2014). Many urban areas are situated in coastal

123
J. M. Edwards et al.

locations or close to significant orography, and these factors increase the complexity of mod-
elling (Fernando 2010).
On coarser scales, such as those of current global climate models, it may suffice to use
a traditional surface scheme, but with modified parameters; at higher resolution, and when
the focus is on conditions in urban canyons, more realistic and detailed parametrizations
are appropriate (Masson 2006; Grimmond et al. 2011). Urban canopy models that explicitly
account for the physical processes of the urban environment are under development (e.g.
Martilli et al. 2002), though the problem of specifying appropriate urban parameters for
more elaborate schemes may make them difficult to use (Grimmond et al. 2011). However,
Ronda et al. (2017) showed that weather forecasting at intra-urban scales can be achieved by
exploiting recent advances in topographic element mapping and aerial photography, as well
as considering detailed mappings of soil characteristics and urban morphological properties.

4.2.5 The Cryosphere

Boundary layers over snow and land or sea-ice in the harsh conditions of the polar regions
constitute some of the most poorly observed and understood manifestations of the PBL, and
the improvement of forecasting capabilities for these regions is seen as a priority for research
(Jung et al. 2016). Climate models exhibit large variations in their simulation of the surface
energy budget of the polar regions (Sorteberg et al. 2007) and unrealistic compensation
between the components of that budget (Boeke and Taylor 2016).
The accurate simulation of the PBL over snow depends not only on the parametrization
of the PBL itself, but also on an accurate representation of the snow pack (Sterk et al. 2013,
2016). Lying snow undergoes complex processes of metamorphism that change its density,
thermal conductivity and albedo (Taillandier et al. 2007; Hachikubo et al. 2014; Wang and
Baker 2014): these processes proceed more rapidly at higher temperatures. The surface energy
budget is therefore strongly influenced by the structure of the snow pack, yet this may not
be simulated well, even by detailed snow-physics models (Domine et al. 2019). Current
NWP and climate models often contain simple snow schemes, but schemes of intermediate
complexity with some representation of snow physics, either with a single snow layer (Dutra
et al. 2010) or with several layers (Dutra et al. 2012; van Kampenhout et al. 2017; Walters
et al. 2019), yield improved performance and are becoming operational.
Whilst the PBL above snow or ice surfaces may often be very stable, it may be very
unstable in cold-air outbreaks, when cold air flows out over open sea (Papritz et al. 2015),
often leading to the development of severe weather, such as polar lows. This transition may
occur very quickly across the marginal ice zone, where drag coefficients are larger than over
solid ice or the open sea. The parametrization of surface exchange in the marginal ice zone is
of great importance. For example, Renfrew et al. (2019) show how the introduction of a more
physically-based representation of form drag in the marginal ice zone significantly improves
the simulation of winds and temperatures in the Unified Model.

4.2.6 The Ocean Surface

Coupled atmosphere–ocean models are required for climate, seasonal and medium-range
weather forecasting, but short-range forecasting systems have generally been uncoupled.
However, the benefits of coupling even for these forecast ranges are becoming appreciated, for
example in the case of tropical cyclones that significantly affect the sea-surface temperature
(SST) (Sandery et al. 2010). More generally, the relation between boundary-layer structure

123
Representation of Boundary-Layer Processes in Numerical…

and SST is not fully understood. Although the ocean surface is fairly homogeneous, some
heterogeneity effects have been reported in relation to SST gradients. Particularly in areas
with sharp SST gradients, e.g., at the edge of the cold tongue in the tropical East Pacific,
advection from cold to warm or from warm to cold leads to an increase or decrease of
near-surface wind speed respectively (Chelton 2005). Atmospheric models underestimate
this modulation of wind speed, most likely because of deficiencies in the turbulent diffusion
schemes. Also, lack of horizontal resolution in models and the SST field plays a role. This
feature has similarities with the diurnal cycle of wind over land, which is modulated by
stability and underestimated in many models (Holtslag et al. 2013). How important it is for
atmosphere-ocean coupling is not yet clear.
At higher resolution, when tropical cyclones are well resolved, the intensity of the cyclone
depends critically on air–sea interaction at high wind speeds. Emanuel (1995) presents a ther-
modynamic argument for the ratio of momentum and scalar transfer coefficients in tropical
cyclones. He concludes that, for rapid intensification and wind speed above 30 m s−1 , it is crit-
ical that the drag coefficient decrease with increased wind speed. With increasing horizontal
resolution, this regime is becoming relevant even for global NWP models (Fig. 1a) Donelan
(2018) presents results from momentum-budget analyses and laboratory wave-channel exper-
iments. He proposes that the drag coefficient increases from about 0.001 at low wind speeds
to nearly 0.003 at about 30 m s−1 and subsequently decreases to 0.0015 at 50 m s−1 . Evi-
dence from modelling centres indeed confirms the benefit for the simulated wind speed of
decreasing drag coefficients at hurricane wind speeds, although the formulation can be rather
ad hoc, e.g., by an empirical reduction of the aerodynamic roughness length (Fan et al. 2012;
Du et al. 2017). There is still some controversy concerning the physical mechanisms. Spray
is an obvious factor that invalidates MOST, but also flow separation and “wave valley filling”
effects have been suggested (Donelan 2018). It can be concluded that the uncertainties in
transfer coefficients for extreme wind speeds over the ocean are still very large (Bell et al.
2012).

5 Summary

Parametrizations of the PBL have developed greatly over the past fifty years, building upon
increased understanding of key physical processes derived from LES and observations. In
contrast to the earliest simple diffusive schemes, current parametrizations explicitly account
for processes such as non-gradient transports, entrainment, and the formation of cloud: these
are crucial for the simulation of stratocumulus-capped boundary layers (Köhler et al. 2011).
Increases in model resolution continue to lead to more accurate and detailed forecasts,
while there is a growing awareness of the importance of interactions between different phys-
ical processes for both NWP and climate modelling. In response to the challenges raised
by these developments, research on the parametrization of the PBL remains essential to the
further improvement of numerical models.

Acknowledgements We thank Ian Boutle for carrying out the simulation used in Fig. 1c and d. and Wayne
Angevine for comments on an earlier draft. An anonymous reviewer’s suggestions were very helpful in
enhancing the breadth of the discussion.

123
J. M. Edwards et al.

References
Albrecht BA, Randall DA, Nicholls S (1988) Observations of marine stratocumulus clouds during FIRE.
Bull Am Meteorol Soc 69:618–626. https://doi.org/10.1175/1520-0477(1988)069<0618:OOMSCD>2.
0.CO;2
Ayotte KW, Sullivan PP, Andrén A, Doney SC, Holtslag AAM, Large WG, McWilliams JC, Moeng CH, Otte
MJ, Tribbia JJ, Wyngaard JC (1996) An evaluation of neutral and convective planetary boundary-layer
parameterizations relative to large-eddy simulations. Boundary-Layer Meteorol 79:131–175. https://doi.
org/10.1007/BF00120078
Baklanov A, Grimmond CSB, Carlson D, Terblanche D, Tang X, Bouchet V, Lee B, Langendijk G, Kolli
RK, Hovsepyan A (2018) From urban meteorology, climate and environment research to integrated city
services. Urban Clim 23:330–341. https://doi.org/10.1016/j.uclim.2017.05.004
Baldauf M, Seifert A, Förstner J, Majewski D, Raschendorfer M, Reinhardt T (2011) Operational convective-
scale numerical weather prediction with the COSMO model: description and sensitivities. Mon Weather
Rev 139:3887–3905. https://doi.org/10.1175/MWR-D-10-05013.1
Balsamo G, Agusti-Panareda A, Albergel C, Arduini G, Beljaars A, Bidlot J, Blyth E, Bousserez N, Boussetta
S, Brown A, Choulga M, Cloke H, Cronin MF, Dahoui M, De Rosnay P, Dirmeyer PA, Drusch M,
Dutra E, Ek MB, Gentine P, Hewitt H, Keeley SP, Kerr Y, Kumar S, Lupu C, Mahfouf JF, McNorton J,
Mecklenburg S, Mogensen K, Muñoz-Sabater J, Orth R, Rabier F, Reichle R, Ruston B, Pappenberger
F, Sandu I, Seneviratne SI, Tietsche S, Trigo IF, Uijlenhoet R, Wedi N, Woolway RI, Zeng X (2018)
Correction: Balsamo, G., et al. Satellite and in situ observations for advancing global Earth surface
modelling: a review. Remote Sens 11(8):941. https://doi.org/10.3390/rs11080941
Barlow JF (2014) Progress in observing and modelling the urban boundary layer. Urban Clim 10:216–240.
https://doi.org/10.1016/j.uclim.2014.03.011
Basu S, Porté-Agel F, Foufoula-Georgiou E, Vinuesa JF, Pahlow M (2006) Revisiting the local scaling hypoth-
esis in stably stratified atmospheric boundary-layer turbulence: an integration of field and laboratory
measurements with large-eddy simulations. Boundary-Layer Meteorol 119(3):473–500. https://doi.org/
10.1007/s10546-005-9036-2
Bauer P, Thorpe A, Brunet G (2015) The quiet revolution of numerical weather prediction. Nature 525:47–55.
https://doi.org/10.1038/nature14956
Bechtold P, Cuijpers J, Mascart P, Trouilhet P (1995) Modeling of trade wind cumuli with a low-order turbulence
model: toward a unified description of Cu and Se clouds in meteorological models. J Atmos Sci 52(4):455–
463. https://doi.org/10.1175/1520-0469(1995)052<0455:MOTWCW>2.0.CO;2
Beljaars ACM (1982) The derivation of fluxes from profiles in perturbed areas. Boundary-Layer Meteorol
24(1):35–55. https://doi.org/10.1007/BF00121798
Beljaars A (1991) Numerical schemes for parameterizations. In: Seminar on numerical methods in atmo-
spheric models, 9–13 Sept 1991, ECMWF, Shinfield Park, Reading, pp 1–42. https://www.ecmwf.int/
sites/default/files/elibrary/1991/8028-numerical-schemesparametrizations.pdf
Beljaars ACM (1995) The parametrization of surface fluxes in large-scale models under free convection. Q J
R Meteorol Soc 121(522):255–270. https://doi.org/10.1002/qj.49712152203
Beljaars A, Betts A (1992) Validation of the boundary layer representation in the ECMWF model. In: Seminar
on validation of models over Europe, 7–11 Sept 1992, ECMWF, Shinfield Park, Reading, vol II, pp
159–196. https://www.ecmwf.int/node/8030
Beljaars ACM, Holtslag AAM (1991) Flux parameterization over land surfaces for atmospheric models. J
Appl Meteorol 30(3):327–341. https://doi.org/10.1175/1520-0450(1991)030<0327:FPOLSF>2.0.CO;
2
Beljaars ACM, Viterbo P (1998) Role of the boundary layer in a numerical weather prediction model. In:
Holtslag AAM, Duynkerke P (eds) Clear and cloudy boundary layers. Royal Netherlands Academy of
Arts and Sciences, Amsterdam, pp 287–304
Beljaars ACM, Schotanus P, Nieuwstadt FTM (1983) Surface layer similarity under nonuniform fetch con-
ditions. J Clim Appl Meteorol 22(10):1800–1810. https://doi.org/10.1175/1520-0450(1983)022<1800:
SLSUNF>2.0.CO;2
Bell MM, Montgomery MT, Emanuel KA (2012) Air-sea enthalpy and momentum exchange at major hurricane
wind speeds observed during CBLAST. J Atmos Sci 69(11):3197–3222. https://doi.org/10.1175/JAS-
D-11-0276.1
Benoit R, Côté J, Mailhot J (1989) Inclusion of a TKE boundary layer parameterization in the Cana-
dian regional finite-element model. Mon Weather Rev 117:1669–1694. https://doi.org/10.1175/1520-
0493(1989)117<1726:IOATBL>2.0.CO;2

123
Representation of Boundary-Layer Processes in Numerical…

Berg LK, Lamb PJ (2016) Surface properties and interactions: coupling the land and atmosphere
within the ARM program. Meteorological Monographs 57:23.1–23.17. https://doi.org/10.1175/
AMSMONOGRAPHS-D-15-0044.1
Bergot T, Terradellas E, Cuxart J, Mira A, Liechti O, Mueller M, Nielsen NW (2007) Intercomparison of
single-column numerical models for the prediction of radiation fog. J Appl Meteorol Clim 46:504–521.
https://doi.org/10.1175/JAM2475.1
Blackadar AK (1962) The vertical distribution of wind and turbulent exchange in a neutral atmosphere. J
Geophys Res 67:3095–3102. https://doi.org/10.1029/JZ067i008p03095
Boeke RC, Taylor PC (2016) Evaluation of the Arctic surface radiation budget in CMIP5 models. J Geophys
Res 121:8525–8548. https://doi.org/10.1002/2016JD025099
Bogenschutz PA, Krueger SK (2013) A simplified PDF parameterization of subgrid-scale clouds and turbulence
for cloud-resolving models. J Adv Model Earth Syst 5(2):195–211. https://doi.org/10.1002/jame.20018
Bolin B (1955) Numerical forecasting with the barotropic model. Tellus 7:27–49. https://doi.org/10.3402/
tellusa.v7i1.8770
Bony S, Dufresne JL (2005) Marine boundary layer clouds at the heart of tropical cloud feedback uncertainties
in climate models. Geophys Res Lett 32:L20806. https://doi.org/10.1029/2005GL023851
Bosveld FC, Baas P, Beljaars ACM, Holtslag AAM, Vilá-Guerau J, de Arellano, van de Wiel BJH (2020)
Fifty years of atmospheric boundary-layer research at cabauw serving weather, air quality and climate.
Boundary-Layer Meteorol. https://doi.org/10.1007/s10546-020-00541-w
Bougeault P, Lacarrère P (1989) Parameterization of orography-induced turbulence in a mesobeta-scale model.
Mon Weather Rev 117:1872–1890. https://doi.org/10.1175/1520-0493(1989)117<1872:POOITI>2.0.
CO;2
Boutle IA, Eyre JEJ, Lock AP (2014) Seamless stratocumulus simulation across the turbulent gray zone. Mon
Weather Rev 142:1655–1668. https://doi.org/10.1175/MWR-D-13-00229.1
Boutle IA, Finnenkoetter A, Lock AP, Wells H (2016) The London model: forecasting fog at 333 m resolution.
Q J R Meteorol Soc 142:360–371. https://doi.org/10.1002/qj.2656
Bretherton CS, Wyant MC (1997) Moisture transport, lower-tropospheric stability, and decoupling of cloud-
topped boundary layers. J Atmos Sci 54:148–167. https://doi.org/10.1175/1520-0469(1997)054<0148:
MTLTSA>2.0.CO;2
Brown AR (1996) Large-eddy simulation and parametrization of the baroclinic boundary-layer. Boundary-
Layer Meteorol 122:1779–1798. https://doi.org/10.1002/qj.49712253603
Brown AR, Beare RJ, Edwards JM, Lock AP, Keogh SJ, Milton SF, Walters DN (2008) Upgrades to the
boundary-layer scheme in the Met Office numerical weather prediction model. Boundary-Layer Meteorol
128:117–132. https://doi.org/10.1007/s10546-008-9275-0
Brunke MA, Fairall CW, Zeng X, Eymard L, Curry JA (2003) Which bulk aerodynamic algorithms are least
problematic in computing ocean surface turbulent fluxes? J Clim 16(4):619–635. https://doi.org/10.1175/
1520-0442(2003)016<0619:WBAAAL>2.0.CO;2
Cava D, Giostra U, Katul G (2015) Characteristics of gravity waves over an Antarctic ice sheet during an
austral summer. Atmosphere 6:1271–1289. https://doi.org/10.3390/atmos6091271
Charnock H (1955) Wind stress on a water surface. Q J R Meteorol Soc 81(350):639–640. https://doi.org/10.
1002/qj.49708135027
Chelton DB (2005) The impact of SST specification on ECMWF surface wind stress fields in the eastern
tropical Pacific. J Clim 18(4):530–550. https://doi.org/10.1175/JCLI-3275.1
Cheng A, Xu KM (2006) Simulation of shallow cumuli and their transition to deep convective clouds by cloud-
resolving models with different third-order turbulence closures. Q J R Meteorol Soc 132(615):359–382.
https://doi.org/10.1256/qj.05.29
Cheng Y, Canuto VM, Howard AM (2002) An improved model for the turbulent PBL. J Atmos Sci 59:1550–
1565. https://doi.org/10.1175/1520-0469(2002)059<1550:AIMFTT>2.0.CO;2
Ching J, Rotunno R, LeMone M, Martilli A, Kosovic B, Jimenez PA, Dudhia J (2014) Convectively induced
secondary circulations in fine-grid mesoscale numerical weather prediction models. Mon Weather Rev
142:3284–3302. https://doi.org/10.1175/MWR-D-13-00318.1
Chung D, Matheou G (2014) Large-eddy simulation of stratified turbulence. Part I: a vortex-based subgrid-scale
model. J Atmos Sci 71:1863–1879. https://doi.org/10.1175/JAS-D-13-0126.1
Clarke RH (1970) Recommended methods for the treatment of the boundary layer in numerical models. Aust
Meteorol Mag 18:51–73
Conzemius RJ, Fedorovich E (2006) Dynamics of sheared convective boundary layer entrainment. Part II:
evaluation of bulk model predictions of entrainment flux. J Atmos Sci 63:1179–1199. https://doi.org/10.
1175/JAS3696.1
Copernicus Climate Change Service (2017) ERA5: Fifth generation of ECMWF atmospheric reanalyses of
the global climate. https://cds.climate.copernicus.eu/cdsapp#!/home. Accessed 14 Nov 2019

123
J. M. Edwards et al.

Couvreux F, Hourdin F, Rio C (2010) Resolved versus parametrized boundary-layer plumes. Part I: a
parametrization-oriented conditional sampling in large-eddy simulations. Boundary-Layer Meteorol
134:441–458. https://doi.org/10.1007/s10546-009-9456-5
Couvreux F, Bazile E, Rodier Q, Maronga B, Matheou G, Chinita MJ, Edwards J, Van Stratum BJH, Van
Heerwaarden CC, Huang J, Moene AF, Cheng A, Fuka V, Basu S, Bou-Zeid E, Canut G, Vignon E
(2020) Intercomparison of large-Eddy simulations of the antarctic boundary layer challenged by very
stable stratification. Boundary-Layer Meteorol. https://doi.org/10.1007/s10546-020-00539-4
Cuxart J, Bougeault P, Redelsperger JL (2000) A turbulence scheme allowing for mesoscale and large-eddy
simulations. Q J R Meteorol Soc 126:1–30. https://doi.org/10.1002/qj.49712656202
Cuxart J, Holtslag AAM, Beare RJ, Bazile E, Beljaars A, Cheng A, Conangla L, Ek M, Freedman F, Hamdi
R, Kerstein A, Kitagawa H, Lenderink G, Lewellen D, Mailhot J, Mauritsen T, Perov V, Schayes G,
Steeneveld GJ, Svensson G, Taylor P, Weng W, Wunsch S, Xu KM (2006) Single-column model inter-
comparison for a stably stratified atmospheric boundary layer. Boundary-Layer Meteorol 118:273–303.
https://doi.org/10.1007/s10546-005-3780-1
Cuxart J, Wrenger B, Matjacic B, Mahrt L (2019) Spatial variability of the lower atmospheric boundary layer
over hilly terrain as observed with an RPAS. Atmosphere 10:715. https://doi.org/10.3390/atmos10110715
Deardorff JW (1966) The counter-gradient heat flux in the lower atmosphere and in the laboratory. J Atmos
Sci 23:503–506. https://doi.org/10.1175/1520-0469(1966)023<0503:TCGHFI>2.0.CO;2
Deardorff JW (1972) Parameterization of the planetary boundary layer for use in general circulation models.
Mon Weather Rev 100:93–106. https://doi.org/10.1175/1520-0493(1972)100<0093:POTPBL>2.3.CO;
2
Derbyshire SH (1990) Nieuwstadt’s stable boundary layer revisited. Q J R Meteorol Soc 116(491):127–158.
https://doi.org/10.1002/qj.49711649106
Domine F, Picard G, Morin S, Barrere M, Madore JB, Langlois A (2019) Major issues in simulating some
Arctic snowpack properties using current detailed snow physics models: consequences for the thermal
regime and water budget of permafrost. J Adv Model Earth Syst 11:34–44. https://doi.org/10.1029/
2018MS001445
Donat MG, Pitman AJ, Seneviratne SI (2017) Regional warming of hot extremes accelerated by surface energy
fluxes. Geophys Res Lett 44:7011–7019. https://doi.org/10.1002/2017GL073733
Donelan MA (2018) On the decrease of the oceanic drag coefficient in high winds. J Geophys Res 123(2):1485–
1501. https://doi.org/10.1002/2017JC013394
Dorman JL, Sellers PJ (1989) A global climatology of albedo, roughness length and stomatal resistance for
atmospheric general circulation models as represented by the simple biosphere model (SiB). J Appl
Meteorol 28(9):833–855. https://doi.org/10.1175/1520-0450(1989)028<0833:AGCOAR>2.0.CO;2
Du J, Bolaños R, Guo Larsén X (2017) The use of a wave boundary layer model in SWAN. J Geophys Res
122(1):42–62. https://doi.org/10.1002/2016JC012104
Dutra E, Balsamo G, Viterbo P, Miranda PM, Beljaars A, Schär C, Elder K (2010) An improved snow scheme
for the ECMWF land surface model: description and offline validation. J Hydrometeorol 11:899–916.
https://doi.org/10.1175/2010JHM1249.1
Dutra E, Viterbo P, Miranda PMA, Balsamo G (2012) Complexity of snow schemes in a climate model and its
impact on surface energy and hydrology. J Hydrometeorol 13:521–538. https://doi.org/10.1175/JHM-
D-11-072.1
Duynkerke PG, de Roode SR, van Zanten MC, Calvo J, Cuxart J, Cheinet S, Chlond A, Grenier H, Jonker
PJ, Köhler M et al (2004) Observations and numerical simulations of the diurnal cycle of the EUROCS
stratocumulus case. Q J R Meteorol Soc 130(604):3269–3296. https://doi.org/10.1256/qj.03.139
Emanuel KA (1995) Sensitivity of tropical cyclones to surface exchange coefficients and a revised steady-
state model incorporating eye dynamics. J Atmos Sci 52(22):3969–3976. https://doi.org/10.1175/1520-
0469(1995)052<3969:SOTCTS>2.0.CO;2
Fairall CW, Bradley EF, Hare JE, Grachev AA, Edson JB (2003) Bulk parameterization of air-sea fluxes:
Updates and verification for the COARE algorithm. J Clim 16(4):571–591. https://doi.org/10.1175/
1520-0442(2003)016<0571:BPOASF>2.0.CO;2
Fan Y, Lin SJ, Held IM, Yu Z, Tolman HL (2012) Global ocean surface wave simulation using a coupled
atmosphere-wave model. J Clim 25(18):6233–6252. https://doi.org/10.1175/JCLI-D-11-00621.1
Fernando HJS (2010) Fluid dynamics of urban atmospheres in complex terrain. Annu Rev Fluid Mech 42:365–
389. https://doi.org/10.1146/annurev-fluid-121108-145459
Finnigan JJ (2000) Turbulence in plant canopies. Annu Rev Fluid Mech 32(1):519–571. https://doi.org/10.
1146/annurev.fluid.32.1.519
Frech M, Mahrt L (1995) A two-scale mixing formulation for the atmospheric boundary layer. Boundary-Layer
Meteorol 73:91–104. https://doi.org/10.1007/BF00708931

123
Representation of Boundary-Layer Processes in Numerical…

Garratt JR (1993) Sensitivity of climate simulations to land-surface and atmospheric boundary-


layer treatments—a review. J Clim 6:419–448. https://doi.org/10.1175/1520-0442(1993)006<0419:
SOCSTL>2.0.CO;2
Garratt JR (1994) The atmospheric boundary layer. Cambridge atmospheric and space science series. Cam-
bridge University Press, Cambridge
Geleyn JF (1988) Interpolation of wind, temperature and humidity values from model levels to the height of
measurement. Tellus A 40(4):347–351. https://doi.org/10.3402/tellusa.v40i4.11805
Genthon C, Six D, Gallée H, Grigioni P, Pellegrini A (2013) Two years of atmospheric boundary layer
observations on a 45-m tower at Dome C on the Antarctic plateau. J Geophys Res 118:3218–3232.
https://doi.org/10.1002/jgrd.50128
Giorgetta MA, Brokopf R, Crueger T, Esch M, Fiedler S, Helmert J, Hohenegger C, Kornblueh L, Köhler
M, Manzini E, Mauritsen T, Nam C, Raddatz T, Rast S, Reinert D, Sakradzija M, Schmidt H, Schneck
R, Schnur R, Silvers L, Wan H, Zängl G, Stevens B (2018) ICON-A, the atmosphere component of the
ICON Earth system model: I. Model description. J Adv Model Earth Syst 10:1613–1637. https://doi.org/
10.1029/2017MS001242
Goger B, Rotach MW, Gohm A, Fuhrer O, Stiperski I, Holtslag AAM (2018) The impact of three-dimensional
effects on the simulation of turbulence kinetic energy in a major Alpine valley. Boundary-Layer Meteorol
168:1–27. https://doi.org/10.1007/s10546-018-0341-y
Golaz JC, Larson VE, Cotton WR (2002a) A PDF-based model for boundary layer clouds. Part I: method and
model description. J Atmos Sci 59(24):3540–3551. https://doi.org/10.1175/1520-0469(2002)059<3540:
APBMFB>2.0.CO;2
Golaz JC, Larson VE, Cotton WR (2002b) A PDF-based model for boundary layer clouds. Part II: model results.
J Atmos Sci 59(24):3552–3571. https://doi.org/10.1175/1520-0469(2002)059<3552:APBMFB>2.0.
CO;2
Golaz JC, Larson VE, Hansen JA, Schanen DP, Griffin BM (2007) Elucidating model inadequacies in a cloud
parameterization by use of an ensemble-based calibration framework. Mon Weather Rev 135(12):4077–
4096. https://doi.org/10.1175/2007MWR2008.1
Grachev AA, Fairall CW, Persson POG, Andreas EL, Guest PS (2005) Stable boundary-layer scaling regimes:
the Sheba data. Boundary-Layer Meteorol 116:201–235. https://doi.org/10.1007/s10546-004-2729-0
Grant ALM, Mason PJ (1990) Observations of boundary-layer structure over complex terrain. Q J R Meteorol
Soc 116(491):159–186. https://doi.org/10.1002/qj.49711649107
Grenier H, Bretherton CS (2001) A moist PBL parameterization for large-scale models and its application to
subtropical cloud-topped marine boundary layers. Mon Weather Rev 129:357–377. https://doi.org/10.
1175/1520-0493(2001)129<0357:AMPPFL>2.0.CO;2
Grimmond CSB, Blackett M, Best MJ, Baik JJ, Belcher SE, Beringer J, Bohnenstengel SI, Calmet I, Chen F,
Coutts A, Dandou A, Fortuniak K, Gouvea ML, Hamdi R, Hendry M, Kanda M, Kawai T, Kawamoto Y,
Kondo H, Krayenhoff ES, Lee SH, Loridan T, Martilli A, Masson V, Miao S, Oleson K, Ooka R, Pigeon
G, Porson A, Ryu YH, Salamanca F, Steeneveld GJ, Tombrou M, Voogt JA, Young DT, Zhang N (2011)
Initial results from Phase 2 of the international urban energy balance model comparison. Int J Climatol
31:244–272. https://doi.org/10.1002/joc.2227
Guo H, Golaz JC, Donner L, Wyman B, Zhao M, Ginoux P (2015) CLUBB as a unified cloud parame-
terization: opportunities and challenges. Geophys Res Lett 42(11):4540–4547. https://doi.org/10.1002/
2015GL063672
Hachikubo A, Yamaguchi A, Arakawa H, Tanikawa T (2014) Effects of temperature and grain type on time
variation of snow specific surface area. Bull Glaciol Res 32:47–53. https://doi.org/10.5331/bgr.32.47
Han J, Witek ML, Teixeira J, Sun R, Pan H, Fletcher JK, Bretherton CS (2016) Implementation in the NCEP
GFS of a hybrid eddy-diffusivity mass-flux (EDMF) boundary layer parameterization with dissipative
heating and modified stable boundary layer mixing. Weather Forecast 31:341–352. https://doi.org/10.
1175/WAF-D-15-0053.1
Harman IN, Finnigan JJ (2007) A simple unified theory for flow in the canopy and roughness sublayer.
Boundary-Layer Meteorol 123:339–363. https://doi.org/10.1007/s10546-006-9145-6
Harman IN, Finnigan JJ (2008) Scalar concentration profiles in the canopy and roughness sublayer. Boundary-
Layer Meteorol 129:323–351. https://doi.org/10.1007/s10546-008-9328-4
Högström U (1996) Review of some basic characteristics of the atmospheric surface layer. Boundary-Layer
Meteorol 78:215–246. https://doi.org/10.1007/BF00120937
Holtslag AAM (1984) Estimates of diabatic wind speed profiles from near-surface weather observations.
Boundary-Layer Meteorol 29:225–250. https://doi.org/10.1007/BF00119790
Holtslag AAM, Boville BA (1993) Local versus nonlocal boundary-layer diffusion in a global climate model.
J Clim 6:1825–1842. https://doi.org/10.1175/1520-0442(1993)006<1825:LVNBLD>2.0.CO;2

123
J. M. Edwards et al.

Holtslag AAM, Moeng CH (1991) Eddy diffusivity and countergradient transport in the convective atmo-
spheric boundary layer. J Atmos Sci 48:1690–1698. https://doi.org/10.1175/1520-0469(1991)048<1690:
EDACTI>2.0.CO;2
Holtslag AAM, Svensson G, Baas P, Basu S, Beare B, Beljaars ACM, Bosveld FC, Cuxart J, Lindvall J,
Steeneveld GJ, Tjernström M, van de Wiel BJH (2013) Stable atmospheric boundary layers and diurnal
cycles: challenges for weather and climate models. Bull Am Meteorol Soc 94(11):1691–1706. https://
doi.org/10.1175/BAMS-D-11-00187.1
Hong SY, Pan HL (1996) Nonlocal boundary layer vertical diffusion in a medium-range forecast model. Mon
Weather Rev 124:2322–2339. https://doi.org/10.1175/1520-0493(1996)124<2322:NBLVDI>2.0.CO;2
Hong SY, Noh Y, Dudhia J (2006) A new vertical diffusion package with an explicit treatment of entrainment
processes. Mon Weather Rev 134:2318–2341. https://doi.org/10.1175/MWR3199.1
Honnert R (2016) Representation of the grey zone of turbulence in the atmospheric boundary layer. Adv Sci
Res 13:63–67. https://doi.org/10.5194/asr-13-63-2016
Honnert R, Masson V, Couvreux F (2011) A diagnostic for evaluating the representation of turbulence in
atmospheric models at the kilometric scale. J Atmos Sci 68:3112–3131. https://doi.org/10.1175/JAS-D-
11-061.1
Hu X, Nielsen-Gammon JW, Zhang F (2010) Evaluation of three planetary boundary layer schemes in the
WRF model. J Appl Meteorol Clim 49:1831–1844. https://doi.org/10.1175/2010JAMC2432.1
Ito J, Niino H, Nakanishi M, Moeng CH (2015) An extension of the Mellor-Yamada model to the Terra
Incognita zone for dry convective mixed layers in the free convection regime. Boundary-Layer Meteorol
157:23–43. https://doi.org/10.1007/s10546-015-0045-5
Janssen PAEM, Doyle JD, Bidlot J, Hansen B, Isaksen L, Viterbo P (2002) Impact and feedback of ocean
waves on the atmosphere. In: Perrie W (ed) Atmosphere–Ocean interactions volume 1, Advances in fluid
mechanics, 33, Computational Mechanics Publications–WIT Press, pp 155–198
Jung T, Gordon ND, Bauer P, Bromwich DH, Chevallier M, Day JJ, Dawson J, Doblas-Reyes F, Fairall C,
Goessling HF, Holland M, Inoue J, Iversen T, Klebe S, Lemke P, Losch M, Makshtas A, Mills B, Nurmi
P, Perovich D, Reid P, Renfrew IA, Smith G, Svensson G, Tolstykh M, Yang Q (2016) Advancing polar
prediction capabilities on daily to seasonal time scales. Bull Am Meteorol Soc 97:1631–1647. https://
doi.org/10.1175/BAMS-D-14-00246.1
Kaimal JC, Wyngaard JC (1990) The Kansas and Minnesota experiments. Boundary-Layer Meteorol 50:31–47.
https://doi.org/10.1007/BF00120517
Kealey JC, Efstathiou GA, Beare RJ (2019) The onset of resolved boundary-layer turbulence at grey-zone
resolutions. Boundary-Layer Meteorol 171:31–52. https://doi.org/10.1007/s10546-018-0420-0
Köhler M, Ahlgrimm M, Beljaars A (2011) Unified treatment of dry convective and stratocumulus-topped
boundary layers in the ECMWF model. Q J R Meteorol Soc 137:43–57. https://doi.org/10.1002/qj.713
Large WG, Pond S (1982) Sensible and latent heat flux measurements over the ocean. J Phys Oceanogr
12(5):464–482. https://doi.org/10.1175/1520-0485(1982)012<0464:SALHFM>2.0.CO;2
Larson VE, Golaz JC (2005) Using probability density functions to derive consistent closure relation-
ships among higher-order moments. Mon Weather Rev 133(4):1023–1042. https://doi.org/10.1175/
MWR2902.1
Larson VE, Schanen DP, Wang M, Ovchinnikov M, Ghan S (2012) PDF parameterization of boundary layer
clouds in models with horizontal grid spacings from 2 to 16 km. Mon Weather Res 140(1):285–306.
https://doi.org/10.1175/MWR-D-10-05059.1
Lawrence BN, Rezny M, Budich R, Bauer P, Behrens J, Carter M, Deconinck W, Ford R, Maynard C,
Mullerworth S, Osuna C, Porter A, Serradell K, Valcke S, Wedi N, Wilson S (2018) Crossing the
chasm: How to develop weather and climate models for next generation computers? Geosci Model Dev
11:1799–1821. https://doi.org/10.5194/gmd-11-1799-2018
Lehner M, Rotach MW (2018) Current challenges in understanding and predicting transport and exchange in
the atmosphere over mountainous terrain. Atmosphere 9:276. https://doi.org/10.3390/atmos9070276
LeMone MA, Pennell WT (1976) The relationship of trade wind cumulus distribution to subcloud layer fluxes
and structure. Mon Weather Rev 104:524–539. https://doi.org/10.1175/1520-0493(1976)104<0524:
TROTWC>2.0.CO;2
LeMone MA, Grossman RL, Coulter RL, Wesley ML, Klazura GE, Poulos GS, Blumen W, Lundquist JK,
Cuenca RH, Kelly SF, Brandes EA, Oncley SP, McMillen RT, Hicks BB (2000) Land-atmosphere interac-
tion research, early results, and opportunities in the Walnut River watershed in southeast Kansas: CASES
and ABLE. Bull Am Meteorol Soc 81:757–780. https://doi.org/10.1175/1520-0477(2000)081<0757:
LIRERA>2.3.CO;2
LeMone MA, Angevine WM, Bretherton CS, Chen F, Dudhia J, Fedorovich E, Katsaros KB, Lenschow
DH, Mahrt L, Patton EG, Sun J, Tjernström M, Weil J (2019) 100 years of progress in boundary-layer

123
Representation of Boundary-Layer Processes in Numerical…

meteorology. Meteorological Monographs. https://doi.org/10.1175/AMSMONOGRAPHS-D-18-0013.


1
Lenderink G, Holtslag AAM (2000) Evaluation of the kinetic energy approach for modeling turbulent fluxes
in stratocumulus. Mon Weather Rev 128:244–258. https://doi.org/10.1175/1520-0493(2000)128<0244:
EOTKEA>2.0.CO;2
Lenderink G, Holtslag AAM (2004) An updated length-scale formulation for turbulent mixing in clear and
cloudy boundary layers. Q J R Meteorol Soc 130:3405–3427. https://doi.org/10.1256/qj.03.117
Lenderink G, VanZanten MC, Duynkerke PG (1999) Can an E-l turbulence closure simulate entrainment in
radiatively driven convective boundary layers? J Atmos Sci 56(18):3331–3337. https://doi.org/10.1175/
1520-0469(1999)056<3331:CAELTC>2.0.CO;2
Li Y, Gao Z, Lenschow DH, Chen F (2010) An improved approach for parameterizing surface-layer turbulent
transfer coefficients in numerical models. Boundary-Layer Meteorol 137(1):153–165. https://doi.org/10.
1007/s10546-010-9523-y
Lilly DK (1968) Models of cloud-topped mixed layers under a strong inversion. Q J R Meteorol Soc 94:292–
309. https://doi.org/10.1002/qj.49709440106
Lock AP (1998) The parametrization of entrainment in cloudy boundary layers. Q J R Meteorol Soc 124:2729–
2753. https://doi.org/10.1002/qj.49712455210
Lock AP (2001) The numerical representation of entrainment in parametrizations of boundary layer tur-
bulent mixing. Mon Weather Rev 129:1148–1163. https://doi.org/10.1175/1520-0493(2001)129<1148:
TNROEI>2.0.CO;2
Lock AP (2004) The sensitivity of a GCM’s marine stratocumulus to cloud-top entrainment. Q J R Meteorol
Soc 130:3323–3338. https://doi.org/10.1256/qj.03.114
Lock AP, Brown AR, Bush MR, Martin GM, Smith RNB (2000) A new boundary layer mixing scheme. Part
I: scheme description and single-column model tests. Mon Weather Rev 128:3187–3199. https://doi.org/
10.1175/1520-0493(2000)128<3187:ANBLMS>2.0.CO;2
Louis JF (1979) A parametric model of vertical eddy fluxes in the atmosphere. Boundary-Layer Meteorol
17:187–202. https://doi.org/10.1007/BF00117978
Lumley JL, Khajeh-Nouri B (1975) Computational modeling of turbulent transport. In: Adv. in geophys.,
vol 18. Elsevier, pp 169–192. https://doi.org/10.1016/S0065-2687(08)60460-4
Lundquist KA, Chow FK, Lundquist JK (2012) An immersed boundary method enabling large-eddy simu-
lations of flow over complex terrain in the WRF model. Mon Weather Rev 140:3936–3955. https://doi.
org/10.1175/MWR-D-11-00311.1
Mahrt L (1987) Grid-averaged surface fluxes. Mon Weather Rev 115(8):1550–1560. https://doi.org/10.1175/
1520-0493(1987)115<1550:GASF>2.0.CO;2
Mahrt L (2008) Bulk formulation of surface fluxes extended to weak-wind stable conditions. Q J R Meteorol
Soc 134(630):1–10. https://doi.org/10.1002/qj.197
Mahrt L, Vickers D (2004) Bulk formulation of the surface heat flux. Boundary-Layer Meteorol 110:357–379.
https://doi.org/10.1023/B:BOUN.0000007244.42320.1e
Mahrt L, Vickers D (2006) Extremely weak mixing in stable conditions. Boundary-Layer Meteorol 119(1):19–
39. https://doi.org/10.1007/s10546-005-9017-5
Mahrt L, Sun J (1995) The subgrid velocity scale in the bulk aerodynamic relationship for spa-
tially averaged scalar fluxes. Mon Weather Rev 123(10):3032–3041. https://doi.org/10.1175/1520-
0493(1995)123<3032:TSVSIT>2.0.CO;2
Martilli A (2007) Current research and future challenges in urban mesoscale modelling. Int J Climatol 27:1909–
1918. https://doi.org/10.1002/joc.1620
Martilli A, Clappier A, Rotach MW (2002) An urban surface exchange parameterisation for mesoscale models.
Boundary-Layer Meteorol 104:261–304. https://doi.org/10.1023/A:1016099921195
Mason P (1991) Boundary-layer parametrization in heterogeneous terrain. In: Workshop on fine-scale mod-
elling and the development of parametrization schemes, 16–18 Sept 1991, ECMWF, Shinfield Park,
Reading, pp 275–288
Mason PJ (1988) The formation of areally-averaged roughness lengths. Q J R Meteorol Soc 114(480):399–420.
https://doi.org/10.1002/qj.49711448007
Mason PJ, Thomson DJ (1992) Stochastic backscatter in large-eddy simulations of boundary layers. J Fluid
Mech 242:51–78. https://doi.org/10.1017/S0022112092002271
Masson V (2006) Urban surface modeling and the meso-scale impact of cities. Theor Appl Climatol 84:35–45.
https://doi.org/10.1007/s00704-005-0142-3
Masson V, Champeaux JL, Chauvin F, Meriguet C, Lacaze R (2003) A global database of land surface
parameters at 1-km resolution in meteorological and climate models. J Clim 16(9):1261–1282. https://
doi.org/10.1175/1520-0442(2003)16<1261:AGDOLS>2.0.CO;2

123
J. M. Edwards et al.

Mauritsen T, Svensson G, Zilitinkevich SS, Esau I, Enger L, Grisogono B (2007) A total turbulent energy
closure model for neutrally and stably stratified atmospheric boundary layers. J Atmos Sci 64:4113–4126.
https://doi.org/10.1175/2007JAS2294.1
McTaggart-Cowan R, Vaillancourt PA, Zadra A, Chamberland S, Charron M, Corvec S, Milbrandt JA, Paquin
D, Ricard Patoine A, Roch M, Separovic L, Yang J (2019) Modernization of atmospheric physics
parameterization in Canadian NWP. J Adv Model Earth Syst 11:3593–3635. https://doi.org/10.1029/
2019MS001781
Mellado JP (2017) Cloud-top entrainment in stratocumulus clouds. Annu Rev Fluid Mech 49:145–169. https://
doi.org/10.1146/annurev-fluid-010816-060231
Mellor GL, Yamada T (1974) A hierarchy of turbulence closure models for planetary boundary layers. J Atmos
Sci 31:1791–1806. https://doi.org/10.1175/1520-0469(1974)031<1791:AHOTCM>2.0.CO;2
Mellor GL, Yamada T (1982) Development of a turbulence closure model for geophysical fluid problems. Rev
Geophys Space Phys 20:851–875. https://doi.org/10.1029/RG020i004p00851
Miller MJ, Beljaars ACM, Palmer TN (1992) The sensitivity of the ECMWF model to the parameter-
ization of evaporation from the tropical oceans. J Clim 5(5):418–434. https://doi.org/10.1175/1520-
0442(1992)005<0418:TSOTEM>2.0.CO;2
Moeng CH (1984) A large-eddy-simulation model for the study of planetary boundary-layer turbulence. J
Atmos Sci 41:2052–2062. https://doi.org/10.1175/1520-0469(1984)041<2052:ALESMF>2.0.CO;2
Moeng CH (2000) Entrainment rate, cloud fraction, and liquid water path of PBL stratocumulus clouds. J
Atmos Sci 57:3627–3643. https://doi.org/10.1175/1520-0469(2000)057<3627:ERCFAL>2.0.CO;2
Nadeau DF, Pardyjak ER, Higgins CW, Parlange MB (2013) Similarity scaling over a steep alpine slope.
Boundary-Layer Meteorol 147:401–419. https://doi.org/10.1007/s10546-012-9787-5
Nakanishi M, Niino H (2009) Development of an improved turbulence closure model for the atmospheric
boundary layer. J Meteorol Soc Jpn 87:895–912. https://doi.org/10.2151/jmsj.87.895
Neumann P, Düben P, Adamidis A, Bauer P, Brück M, Kornblueh D, Klocke D, Stevens B, Wedi N, Biercamp
J (2019) Assessing the scales in numerical weather and climate predictions: Will exascale be the rescue?
Philos Trans R Soc 377A(20180):148. https://doi.org/10.1098/rsta.2018.0148
Nicholls S, Turton JD (1986) An observational study of the structure of stratiform cloud sheets: part II.
Entrainment. Q J R Meteorol Soc 112:461–480. https://doi.org/10.1002/qj.49711247210
Nieuwstadt FTM (1984) The turbulent structure of the stable, nocturnal boundary layer. J Atmos Sci
41(14):2202–2216. https://doi.org/10.1175/1520-0469(1984)041<2202:TTSOTS>2.0.CO;2
Noh Y, Cheon WG, Hong SY, Raasch S (2003) Improvement of the K-profile model for the planetary boundary
layer based on large-eddy simulation data. Boundary-Layer Meteorol 107:401–427. https://doi.org/10.
1023/A:1022146015946
Olson JB, Kenyon JS, Djalalova I, Bianco L, Turner DD, Pichugina Y, Choukulkar A, Toy MD, Brown JM,
Angevine WM, Akish E, Bao J, Jimenez P, Kosovic B, Lundquist KA, Draxl C, Lundquist JK, McCaa
J, McCaffrey K, Lantz K, Long C, Wilczak J, Banta R, Marquis M, Redfern S, Berg LK, Shaw W, Cline
J (2019) Improving wind energy forecasting through numerical weather prediction model development.
Bull Am Meteorol Soc 100:2201–2220. https://doi.org/10.1175/BAMS-D-18-0040.1
Paluch IR, Lenschow DH, Siems S, Kok GL, Schillawski RD, McKeen S (1994) Evolution of the subtropical
marine boundary layer: comparison of soundings over the eastern Pacific from FIRE and HaRP. J Atmos
Sci 51:1465–1479. https://doi.org/10.1175/1520-0469(1994)051<1465:EOTSMB>2.0.CO;2
Papritz L, Pfahl S, Sodemann H, Wernli H (2015) A climatology of cold air outbreaks and their impact on
air-sea heat fluxes in the high-latitude South Pacific. J Clim 28:342–364. https://doi.org/10.1175/JCLI-
D-14-00482.1
Pleim JE (2007) A combined local and nonlocal closure model for the atmospheric boundary layer. Part I:
model description and testing. J Appl Meteorol Clim 46:1383–1395. https://doi.org/10.1175/JAM2539.
1
Polichtchouk I, Shepherd TG (2016) Zonal-mean circulation response to reduced air-sea momentum roughness.
Q J R Meteorol Soc 142(700):2611–2622. https://doi.org/10.1002/qj.2850
Prandtl L (1925) Bericht über Untersuchungen zur ausgebildeten Turbulenz. Z Angew Math Mech 5:136–139.
https://doi.org/10.1002/zamm.19250050212
Prein AF, Gobiet A, Suklitsch M, Truhetz H, Awan NK, Keuler K, Georgievski G (2013) Added value of
convection permitting seasonal simulations. Clim Dyn 41:2655–2677. https://doi.org/10.1007/s00382-
013-1744-6
Price JD, Lane S, Boutle IA, Smith DK, Bergot T, Lac C, Duconge L, McGregor J, Kerr-Munslow A, Pickering
M, Clark R (2018) LANFEX: a field and modeling study to improve our understanding and forecasting
of radiation fog. Bull Am Meteorol Soc 99:2061–2077. https://doi.org/10.1175/BAMS-D-16-0299.1

123
Representation of Boundary-Layer Processes in Numerical…

Randall D, Krueger S, Bretherton C, Curry J, Duynkerke P, Moncrieff M, Ryan B, Starr D, Miller M, Rossow
W, Tselioudis G, Wielicki B (2003) Confronting models with data: the GEWEX cloud systems study.
Bull Am Meteorol Soc 84:455–470. https://doi.org/10.1175/BAMS-84-4-455
Raupach MR, Finnigan JJ (1988) ‘Single-layer models of evaporation from plant canopies are incorrect but
useful, whereas multilayer models are correct but useless’: discuss. Aust J Plant Physiol. https://doi.org/
10.1071/PP9880705
Reigber A, Moreira A (2000) First demonstration of airborne SAR tomography using multibaseline L-band
data. IEEE Trans Geosci Remote 38(5):2142–2152. https://doi.org/10.1109/36.868873
Renfrew IA, Elvidge AD, Edwards JM (2019) Atmospheric sensitivity to marginal-ice-zone drag: local and
global responses. Q J R Meteorol Soc 145:1165–1179. https://doi.org/10.1002/qj.3486
Roberts MJ, Vidale PL, Senior C, Hewitt HT, Bates C, Berthou S, Chang P, Christensen HM, Danilov S,
Demory M, Griffies SM, Haarsma R, Jung T, Martin G, Minobe S, Ringler T, Satoh M, Schiemann
R, Scoccimarro E, Stephens G, Wehner MF (2018) The benefits of global high resolution for climate
simulation: process understanding and the enabling of stakeholder decisions at the regional scale. Bull
Am Meteorol Soc 99:2341–2359. https://doi.org/10.1175/BAMS-D-15-00320.1
Rodier Q, Masson V, Couvreux F, Paci A (2017) Evaluation of a buoyancy and shear based mixing length for
a turbulence scheme. Front Earth Sci 5:65. https://doi.org/10.3389/feart.2017.00065
Ronda RJ, Steeneveld GJ, Heusinkveld BG, Attema JJ, Holtslag AAM (2017) Urban finescale forecasting
reveals weather conditions with unprecedented detail. Bull Am Meteorol Soc 98:2675–2688. https://doi.
org/10.1175/BAMS-D-16-0297.1
Saito K, Ishida J, Aranami K, Hara T, Segawa T, Narita M, Honda Y (2007) Nonhydrostatic atmospheric
models and operational development at JMA. J Meteorol Soc Jpn 85B:271–304. https://doi.org/10.2151/
jmsj.85B.271
Sandery PA, Brassington GB, Craig A, Pugh T (2010) Impacts of ocean-atmosphere coupling on tropical
cyclone intensity and ocean prediction in the Australian region. Mon Weather Rev 138:2074–2091.
https://doi.org/10.1175/2010MWR3101.1
Sandu I, Beljaars A, Balsamo G, Ghelli A (2011) Revision of the surface roughness length table. ECMWF
Newsletter 130:8–10
Sandu I, Beljaars A, Bechtold P, Mauritsen T, Balsamo G (2013) Why is it so difficult to represent stably
stratified conditions in numerical weather prediction (NWP) models. J Adv Model Earth Syst 5:117–
133. https://doi.org/10.1002/jame.20013
Santanello JA, Dirmeyer PA, Ferguson CR, Findell KL, Tawfik AB, Berg A, Ek M, Gentine P, Guillod BP, van
Heerwaarden C, Roundy J, Wulfmeyer V (2018) Land-atmosphere interactions: the LoCo perspective.
Bull Am Meteorol Soc 99:1253–1272. https://doi.org/10.1175/BAMS-D-17-0001.1
Schubert WH (1976) Experiments with Lilly’s cloud-topped mixed layer model. J Atmos Sci 33:436–446.
https://doi.org/10.1175/1520-0469(1976)033<0436:EWLCTM>2.0.CO;2
Shin HH, Hong SY (2015) Representation of the subgrid-scale turbulent transport in convective boundary
layers at gray-zone resolutions. Mon Weather Rev 143:250–271. https://doi.org/10.1175/MWR-D-14-
00116.1
Siebesma AP, Cuipers JWM (1995) Evaluation of parametric assumptions for shallow cumulus convection. J
Atmos Sci 52:650–666. https://doi.org/10.1175/1520-0469(1995)052<0650:EOPAFS>2.0.CO;2
Siebesma AP, Soares PMM, Teixeira J (2007) A combined eddy-diffusivity mass-flux approach for the con-
vective boundary layer. J Atmos Sci 64:1230–1248. https://doi.org/10.1175/JAS3888.1
Sillmann J, Thorarinsdottir T, Keenlyside N, Schaller N, Alexander LV, Hegerl G, Seneviratne SI, Vautard
R, Zhang X, Zwiers FW (2017) Understanding, modeling and predicting weather and climate extremes:
challenges and opportunities. Weather Clim Extremes 18:65–74. https://doi.org/10.1016/j.wace.2017.
10.003
Sisterson DL, Peppler RA, Cress TS, Lamb PJ, Turner DD (2016) The ARM Southern Great Plains (SGP) site.
Meteorological Monographs 57:6.1–6.4. https://doi.org/10.1175/AMSMONOGRAPHS-D-16-0004.1
Smagorinsky J, Manabe S, Leith Holloway Jr J (1965) Numerical results from a nine-level general
circulation model of the atmosphere. Mon Weather Rev 93:727–768. https://doi.org/10.1175/1520-
0493(1965)093<0727:NRFANL>2.3.CO;2
Sorteberg A, Kattsov V, Walsh JE, Pavlova T (2007) The Arctic surface energy budget as simulated with the
IPCC AR4 AOGCMs. Clim Dyn 29:131–156. https://doi.org/10.1007/s00382-006-0222-9
Steeneveld GJ, Nappo CJ, Holtslag AAM (2009) Estimation of orographically induced wave drag in the stable
boundary layer during the CASES-99 experimental campaign. Acta Geophys 57(4):857–881. https://doi.
org/10.2478/s11600-009-0028-3
Steeneveld GJ, Ronda RJ, Holtslag AAM (2015) The challenge of forecasting the onset and development of
radiation fog using mesoscale atmospheric models. Boundary-Layer Meteorol 154:265–289. https://doi.
org/10.1007/s10546-014-9973-8

123
J. M. Edwards et al.

Sterk HAM, Steeneveld GJ, Holtslag AAM (2013) The role of snow-surface coupling, radiation, and turbulent
mixing in modeling a stable boundary layer over Arctic sea ice. J Geophys Res 118:1199–1217. https://
doi.org/10.1002/jgrd.50158
Sterk HAM, Steeneveld GJ, Bosveld FC, Vihma T, Anderson PS, Holtslag AAM (2016) Clear-sky stable
boundary layers with low winds over snow-covered surfaces. Part 2: process sensitivity. Q J R Meteorol
Soc 142:821–835. https://doi.org/10.1002/qj.2684
Stevens B (2000) Quasi-steady analysis of a PBL model with an eddy-diffusivity profile and nonlocal fluxes.
Mon Weather Rev 128:824–836. https://doi.org/10.1175/1520-0493(2000)128<0824:QSAOAP>2.0.
CO;2
Stevens B (2002) Entrainment in stratocumulus-topped mixed layers. Q J R Meteorol Soc 128:2663–2690.
https://doi.org/10.1256/qj.01.202
Stevens B, Moeng C, Sullivan PP (1999) Large-eddy simulations of radiatively driven convection: sensitiv-
ities to the representation of small scales. J Atmos Sci 56:3963–3984. https://doi.org/10.1175/1520-
0469(1999)056<3963:LESORD>2.0.CO;2
Stoll R, Porté-Agel F (2006) Effect of roughness on surface boundary conditions for large-eddy simulation.
Boundary-Layer Meteorol 118(1):169–187. https://doi.org/10.1007/s10546-005-4735-2
Stull RB (1988) An introduction to boundary layer meteorology. Atmospheric and oceanographic sciences
library. Springer, Dordrecht
Stull RB (1991) Static stability—an update. Bull Am Meteorol Soc 72:1521–1529. https://doi.org/10.1175/
1520-0477(1991)072<1521:SSU>2.0.CO;2
Suarez MJ, Arakawa A, Randall DA (1983) The parameterization of the planetary boundary layer in the UCLA
general circulation model: formulation and results. Mon Weather Rev 111:2224–2243. https://doi.org/
10.1175/1520-0493(1983)111<2224:TPOTPB>2.0.CO;2
Sukoriansky S, Galperin B, Perov V (2005) Application of a new spectral theory of stably stratified turbulence
to the atmospheric boundary layer over sea ice. Boundary-Layer Meteorol 117:231–257. https://doi.org/
10.1007/s10546-004-6848-4
Sun J, Lenschow DH, LeMone MA, Mahrt L (2016) The role of large-coherent-eddy transport in the atmo-
spheric surface layer based on CASES-99 observations. Boundary-layer Meteorol 160(1):83–111. https://
doi.org/10.1007/s10546-016-0134-0
Sušelj K, Teixeira J, Chung D (2013) A unified model for moist convective boundary layers based on a
stochastic eddy-diffusivity/mass-flux parameterization. J Atmos Sci 70:1929–1953. https://doi.org/10.
1175/JAS-D-12-0106.1
Taillandier AS, Domine F, Simpson WR, Sturm M, Douglas TA (2007) Rate of decrease of the specific surface
area of dry snow: isothermal and temperature gradient conditions. J Geophys Res 112(F03):003. https://
doi.org/10.1029/2006JF000514
Tang Y, Lean HW, Bornemann J (2013) The benefits of the Met Office variable resolution MWP model for
forecasting convection. Meteorol Appl 20:417–426. https://doi.org/10.1002/met.1300
Tastula EM, Galperin B, Dudhia J, LeMone MA, Sukoriansky S, Vihma T (2014) Methodical assessment of
the differences between the QNSE and MYJ PBL schemes for stable conditions. Q J R Meteorol Soc
141:2077–2089. https://doi.org/10.1002/qj.2503
Taylor KE, Stouffer RJ, Meehl GA (2012) An overview of CMIP5 and the experiment design. Bull Am
Meteorol Soc 93:485–498. https://doi.org/10.1175/BAMS-D-11-00094.1
Tennekes H, Driedonks AGM (1981) Basic entrainment equations for the atmospheric boundary layer.
Boundary-Layer Meteorol 20:515–531. https://doi.org/10.1007/BF00122299
Termonia P, Fischer C, Bazile E, Bouyssel F, Brožková R, Bénard P, Bochenek B, Degrauwe D, Derková M,
Khatib RE, Hamdi R, Mašek J, Pottier P, Pristov N, Seity Y, Smolíková P, Španiel O, Tudor M, Wang
Y, Wittmann C, Joly A (2018) The ALADIN system and its canonical model configurations AROME
CY41T1 and ALARO CY40T1. Geosci Model Dev 11:257–281. https://doi.org/10.5194/gmd-11-257-
2018
Troen IB, Mahrt L (1986) A model of the atmospheric boundary layer; sensitivity to surface evaporation.
Boundary-Layer Meteorol 37:129–148. https://doi.org/10.1007/BF00122760
Tsiringakis A, Steeneveld GJ, Holtslag AAM (2017) Small-scale orographic gravity wave drag in stable
boundary layers and its impact on synoptic systems and near-surface meteorology. Q J R Meteorol Soc
143:1504–1516. https://doi.org/10.1002/qj.3021
Turton JD, Nicholls S (1987) A study of the diurnal variation of stratocumulus using a multiple mixed layer
model. Q J R Meteorol Soc 113:969–1009. https://doi.org/10.1002/qj.49711347712
van Kampenhout L, Lenaerts JTM, Lipscomb WH, Sacks WJ, Lawrence DM, Slater AG, van den Broeke MR
(2017) Improving the representation of polar snow and firn in the community earth system model. J Adv
Model Earth Syst 9:2583–2600. https://doi.org/10.1002/2017MS000988

123
Representation of Boundary-Layer Processes in Numerical…

van Zanten MC, Duynkerke PG, Cuijpers JW (1999) Entrainment parameterization in convective boundary
layers. J Atmos Sci 56:813–828. https://doi.org/10.1175/1520-0469(1999)056<0813:EPICBL>2.0.CO;
2
van der Linden SJA, Edwards JM, van Heerwaarden CC, Vignon E, Genthon C, Petenko I, Baas P, Jonker
HJJ, van de Wiel BJH (2019) Large-eddy simulations of the steady wintertime Antarctic boundary layer.
Boundary-Layer Meteorol 173:165–191. https://doi.org/10.1007/s10546-019-00461-4
Vogelezang DHP, Holtslag AAM (1996) Evaluation and model impacts of alternative boundary-layer height
formulations. Boundary-Layer Meteorol 81:245–269. https://doi.org/10.1007/BF02430331
Vosper S, Carter E, Lean H, Lock A, Clark P, Webster S (2013) High resolution modelling of valley cold pools.
Atmos Sci Lett 14:193–199. https://doi.org/10.1002/asl2.439
Walters D, Baran A, Boutle I, Brooks M, Earnshaw P, Edwards J, Furtado K, Hill P, Lock A, Manners J,
Morcrette C, Mulcahy J, Sanchez C, Smith C, Stratton R, Tennant W, Tomassini L, Van Weverberg
K, Vosper S, Willett M, Browse J, Bushell A, Dalvi M, Essery R, Gedney N, Hardiman S, Johnson
B, Johnson C, Jones A, Mann G, Milton S, Rumbold H, Sellar A, Ujiie M, Whitall M, Williams K,
Zerroukat M (2019) The Met Office Unified Model Global Atmosphere 7.0/7.1 and JULES Global Land
7.0 configurations. Geosci Model Dev 12:1909–1963. https://doi.org/10.5194/gmd-12-1909-2019
Wang X, Baker I (2014) Evolution of the specific surface area of snow during high-temperature gradient
metamorphism. J Geophys Res 119:13,690–13,703. https://doi.org/10.1002/2014JD022131
Webster PJ, Lukas R (1992) TOGA COARE: the coupled ocean-atmosphere response experiment. Bull
Am Meteorol Soc 73(9):1377–1416. https://doi.org/10.1175/1520-0477(1992)073<1377:TCTCOR>2.
0.CO;2
Wedi NP (2014) Increasing horizontal resolution in numerical weather prediction and climate simulations:
Illusion or panacea? Philos Trans R Soc 372A(20130):289. https://doi.org/10.1098/rsta.2013.0289
Wilson K, Goldstein A, Falge E, Aubinet M, Baldocchi D, Berbigier P, Bernhofer C, Ceulemans R, Dolman
H, Field C, Grelle A, Ibrom A, Law BE, Kowalski A, Meyers T, Moncrieff J, Monson R, Oechel W,
Tenhunen J, Valentini R, Verma S (2002) Energy balance closure at FLUXNET sites. Agric For Meteorol
113:223–243. https://doi.org/10.1016/S0168-1923(02)00109-0
Wood N, Mason P (1993) The pressure force induced by neutral, turbulent flow over hills. Q J R Meteorol
Soc 119(514):1233–1267. https://doi.org/10.1002/qj.49711951402
Wulfmeyer V, Baker B, Banta R, Behrendt A, Bonin T, Brewer WA, Buban M, Choukulkar A, Dumas E,
Hardesty RM, Heus T, Ingwersen J, Lange D, Lee TR, Metzendorf S, Muppa SK, Meyers T, Newsom
R, Osman M, Raasch S, Santanello J, Senff C, Spät F, Wagner T, Weckwerth T (2018) A new research
approach for observing and characterizing land-atmosphere feedback. Bull Am Meteorol Soc 99:1639–
1667. https://doi.org/10.1175/BAMS-D-17-0009.1
Wyngaard JC (2004) Toward numerical modelling in the “terra incognita”. J Atmos Sci 61:1816–1826. https://
doi.org/10.1175/1520-0469(2004)061<1816:TNMITT>2.0.CO;2
Xie B, Fung JCH, Chan A, Lau A (2012) Evaluation of nonlocal and local planetary boundary layer schemes
in the WRF model. J Geophys Res 117(D12):103. https://doi.org/10.1029/2011JD017080
Yamada T (1977) A numerical experiment on pollutant dispersion in a horizontally-homogeneous atmospheric
boundary layer. Atmos Environ 11:1015–1024. https://doi.org/10.1016/0004-6981(77)90230-X
Zscheischler J, Westra S, van den Hurk BJJM, Seneviratne SI, Ward PJ, Pitman A, AghaKouchak A, Bresch
DN, Leonard M, Wahl T, Zhang X (2018) Future climate risk from compound events. Nat Clim Change
8:469–477. https://doi.org/10.1038/s41558-018-0156-3

Publisher’s Note Springer Nature remains neutral with regard to jurisdictional claims in published maps and
institutional affiliations.

123

You might also like