You are on page 1of 13

Quarterly Journal of the Royal Meteorological Society Q. J. R. Meteorol. Soc. 143: 1241–1253, April 2017 A DOI:10.1002/qj.

2998

Stable boundary-layer regimes at Dome C, Antarctica: observation


and analysis
Etienne Vignon,a * Bas J. H. van de Wiel,b Ivo G. S. van Hooijdonk,c Christophe Genthon,a
Steven J. A. van der Linden,b J. Antoon van Hooft,b Peter Baas,b William Maurel,d Olivier Traulléd
and Giampietro Casasantae
a Univ. Grenoble Alpes/CNRS/IRD, IGE, F-38000 Grenoble, France
b
Faculty of Civil Engineering and Geosciences, Geoscience and Remote Sensing, Delft University of Technology, Delft, the Netherlands
c
Fluid Dynamics Laboratory and J.M. Burgerscentrum, Eindhoven University of Technology, the Netherlands
d
CNRM-GAME, Toulouse, France
e
Institute of Atmospheric Science and Climate, CNR, Rome, Italy
*Correspondence to: E. Vignon, Institut des Géosciences de l’Environnement, CS 40700 38 058 Grenoble Cedex 9, France. E-mail:
etienne.vignon@univ-grenoble-alpes.fr

Investigation of meteorological measurements along a 45 m tower at Dome C on the high


East Antarctic Plateau revealed two distinct stable boundary layer (SBL) regimes at this
location. The first regime is characterized by strong winds and continuous turbulence. It
results in full vertical coupling of temperature, wind magnitude and wind direction in the
SBL. The second regime is characterized by weak winds, associated with weak turbulent
activity and very strong temperature inversions reaching up to 25 K in the lowest 10 m.
Vertical temperature profiles are generally exponentially shaped (convex) in the first regime
and ‘convex–concave–convex’ in the second.
The transition between the two regimes is particularly abrupt when looking at the
near-surface temperature inversion and it can be identified by a 10 m wind-speed threshold.
With winds under this threshold, the turbulent heat supply toward the surface becomes
significantly lower than the net surface radiative cooling. The threshold value (including its
range of uncertainty) appears to agree with recent theoretical predictions from the so-called
‘minimum wind speed for sustainable turbulence’ (MWST) theory.
For the quasi-steady, clear-sky winter cases, the relation between the near-surface
inversion amplitude and the wind speed takes a characteristic ‘S’ shape. Closer analysis
suggests that this relation corresponds to a ‘critical transition’ between a steady turbulent
and a steady ‘radiative’ regime, with a dynamically unstable branch in the transition zone.
These fascinating characteristics of the Antarctic boundary layer challenge present and
future numerical models to represent this region in a physically correct manner.

Key Words: Antarctic atmosphere; critical transition; stable boundary layer; wind-speed threshold

Received 23 April 2016; Revised 20 December 2016; Accepted 10 January 2017; Published online in Wiley Online Library
27 March 2017

1. Introduction corresponds to cloudy and/or windy conditions associated


with moderate stratification and continuous turbulence. vSBL
Modelling the atmospheric stable boundary layer (hereafter SBL) corresponds to clear-sky and/or weak-wind conditions with
is one of the current major challenges in the climate and weather strong temperature stratification. Mahrt (1998a, 1998b) and more
forecast modelling community today (Holtslag et al., 2013; Mahrt, recently Conangla et al. (2008) further distinguished SBL regimes
2014). To develop and improve parametrizations of the SBL according to the dependence of the surface sensible heat flux on
in atmospheric models requires a better understanding of the z/L, with z the height and L the Monin–Obukhov length. For
physical processes involved and their relative contributions to the wSBL, the sensible heat flux increases with stability because of the
mixing of momentum, heat and moisture. Some methods have increase in surface temperature gradient; for vSBL, the sensible
emerged for classification of the SBL into regimes based on in situ heat flux is damped by thermal stratification and decreases with
observations. A common classification is to distinguish the weakly stability. A similar classification was made for several datasets
stable boundary layer (hereafter wSBL) from the very stable in Mauritsen and Svensson (2007), but using the Richardson
boundary layers (hereafter vSBL). The wSBL regime typically number Ri as stability parameter. The latter study also evidenced


c 2017 Royal Meteorological Society
1242 E. Vignon et al.

the anisotropy of the turbulence and the significant contribution Table 1. Heights of the temperature and wind sensors on 1 January 2015.
of the potential turbulent energy in vSBL.
Furthermore, van de Wiel et al. (2012b) introduced the inno- Reference level height T U and δ
vative concept of the wind-speed threshold Umin , for which the 1m 0.9 m 1.35 m
net cooling of the surface can be balanced by the sensible heat 2m 1.89 m 2.35 m
flux. This threshold, called the ‘minimum wind speed for sus- 3m 2.58 m 3.1 m
tainable turbulence’ (MWST), made it possible to contrast two 10 m 9.94 m 8.62 m
SBL regimes at Cabauw in the Netherlands (van Hooijdonk et 17 m 17.2 m 17.82 m
al., 2015): vSBL for U < Umin and wSBL with U > Umin . Other 25 m 24.66 m 25.18 m
32 m 32.02 m 32.54 m
studies have evidenced wind thresholds that separate SBL regimes.
41 m 41.22 m 40.95 m
Sun et al. (2012) distinguished a strongly turbulent regime from
a weakly turbulent one using a critical wind speed, above which
the root mean square of the turbulent kinetic energy (TKE) and
the friction velocity (u∗ ) increase steadily with an increase in extreme temperature gradients often greater than 1 K m−1 in the
wind speed. Acevedo et al. (2015) further distinguished two SBL first metres above the ground (Genthon et al., 2013; Pietroni et
regimes with√ a ‘crossover’ threshold, defined as the wind velocity al., 2014). The turbulent boundary-layer height does not exceed
a few tens of metres or even a few metres (Pietroni et al., 2012;
at which the TKE and u∗ vertical gradients reverse signs. Above
Casasanta et al., 2014; Petenko et al., 2014).
this threshold, the TKE is generated mainly by mechanical fric-
The present study shows a first-order climatology of the SBL
tion at the surface and propagates upward. Conversely, below the
regimes at Dome C using an extensive meteorological dataset. The
‘crossover’ threshold, the near-surface turbulence is damped due
aim is to foster further in-depth observational analysis, as well
to strong stratification. Kelvin–Helmholz instabilities can occur
as to challenge conceptual and numerical models to reproduce
in the upper part of the boundary layer if the wind shear is strong:
those regimes and the features connected to them. The article
just below a low-level jet peak, for instance (Newsom and Banta,
is structured as follows. In section 2, we briefly describe the
2003; Sun et al., 2012). This can lead to the downward propagation
meteorological dataset of Dome C. In section 3, we present our
of turbulence to the surface, marked by a positive vertical gradi-
method to analyze the SBL regimes. In section 4, we demonstrate
ent of TKE and u∗ , and to a progressive recovery of near-surface
the occurrence of two SBL regimes at Dome C by analyzing
mixing. In vSBL, the near surface continuous turbulence can cut temperature, wind and turbulence behaviour using the 10 m
off and the atmosphere decouples mechanically from the surface wind speed. In section 5, we explore the impact of the radiative
(Banta et al., 2007). This was explained by conceptual models energy supply at the surface on the SBL and in section 6 we
in Derbyshire (1999) and van de Wiel et al. (2007) and using show seasonal differences. In section 7, we discuss theoretical
direct numerical simulations in van Hooijdonk et al. (2017). In predictions from the MWST, the numerical modelling of SBL
such conditions, the radiative cooling of air may even become regimes and an observed ‘S’ shape dependence on the near-
the dominant thermodynamic process (Estournel and Guedalia, surface temperature inversion on the wind speed. In section 8, we
1985), even though some remnants of dynamic mixing of heat present our conclusions.
can persist, like intermittent bursts (e.g van de Wiel et al., 2003)
or internal waves (Zilitinkevich et al., 2008; Sun et al., 2015). 2. In situ data
To evaluate the representation of real atmospheric SBL
regimes, numerical models need to be compared with in situ 2.1. Temperature and wind
measurements. However, the flow in the SBL may depend on the
location, different geographical conditions providing different Dome C is located in the eastern part of the high Antarctic Plateau
forcings for the boundary layer. It is thus interesting to study SBL (75◦ 06 S, 123◦ 20 E, 3233 m above sea level: see e.g. Pietroni et al.
regimes using observations in different places across the globe (2012) for a map of the location). In 2008, a 45 m meteorological
and then to see whether parametrizations of turbulence in models tower was built near the French–Italian station Concordia with
are robust, even in uncommon and extreme boundary layers in six levels of wind and temperature measurements. In 2012,
Antarctica for instance. a 2.5 m mast was added about 300 m from the tower, which
In the eastern part of the Antarctic Plateau, Dome C has samples the wind and temperature very close to the surface.
become a place of particular interest for studying the atmospheric Given the homogeneity of the landscape, in this study we assume
boundary layer for both the atmospheric physics community and vertical continuity of the measurements between the lowest level
astronomers, who require undisturbed telescope observations. of the 2.5 m mast and the highest level of the tower. Table 1
The flatness and homogeneity of the landscape prevent the local gives the exact height of the wind speed (U), wind direction (δ)
generation of katabatic winds by any topographic effects (Aristidi and temperature (T) measurements on 1 January 2015. Changes
et al., 2005). Wind-borne snow can affect the SBL dynamics in in measurement heights due to snow accumulation are taken
coastal Antarctic regions (Gallée et al., 2001; Barral et al., 2014a). into account (Vignon et al., 2017). For the sake of simplicity,
However, as the wind at Dome C is relatively moderate, major in the following, the height of the measurements is indicated
drifting snow events only occur very sporadically (Libois et al., by the reference level height in the first column in Table 1.
2014) and are thus not expected to have a significant impact on the The instruments sample every 30 s, but only half-hourly means,
SBL climatology. Moreover, the boundary layer is subject to a wide maxima, minima and standard deviations were recorded. The
range of stability, from convection to extreme stable stratification. manufacturer specifies that the aerovanes (Young 05013) have a
Indeed, in summer, even though the sun is permanently above 1 m s−1 starting threshold and, even though the consistency of
the horizon, the boundary layer evolves in a typical diurnal cycle wind data were checked visually, wind-speed values below 1 m s−1
similar to that over flat ground at lower latitudes. During the ‘day’, should be interpreted with caution. Detailed information on the
i.e. when the sun is high in the sky, convective activity develops instruments and data processing can be found in Genthon et al.
(Mastrantonio et al., 1999; Argentini et al., 2005; King et al., 2006; (2010, 2013) and Vignon et al. (2017).
Genthon et al., 2010; Ricaud et al., 2012), while during the ‘night’,
i.e. when the sun gets closer to the horizon (although still above), 2.2. Turbulent quantities
the radiative cooling of the surface leads to a stable stratification
(Genthon et al., 2010). Inertial oscillation often produces a low- From 16 December 2009–31 January 2010, four sonic thermo-
level jet at the top of the nocturnal boundary layer, i.e. between 15 anemometers of type Applied Technology Sat 3Sx (hereafter
and 60 m (Gallée et al., 2015). On the other hand, during the polar Atec) were installed at heights of 7, 22, 30 and 37 m on
night in winter, quasi-permanent stable stratification occurs, with the tower. These sensors sample at a frequency of 10 Hz and


c 2017 Royal Meteorological Society Q. J. R. Meteorol. Soc. 143: 1241–1253 (2017)
Stable Boundary Layer Regimes 1243

alternate 8 min measurement intervals with 12 min heating and Svensson (2007) show that the usage of z/L suffers from
periods. Due to the harsh climate conditions, the raw signal self-correlations with other turbulent quantities and prefer using
was quite noisy and the data were consequently processed as Ri. At Dome C, we looked at the dependence of meteorological
follows. variables in the SBL on Ri. Figure 1 shows the dependence of
The signal was de-spiked using a two-step algorithm, inspired the temperature difference from the surface at three different
by Vickers and Mahrt (1997). First, isolated individual spikes heights versus the local Ri. An overall increase of the temperature
Xi for which (Xi − Xi−1 ) and (Xi+1 − Xi ) had opposite signs difference with Ri is visible and this increase is quite monotonic
and having amplitudes greater than a threshold (2.5 m s−1 at heights of 3 and 17 m. One may, however, distinguish two
for horizontal wind components, 0.5 m s−1 for vertical wind ensembles of data: one with weak temperature differences at low
component and 2.5 K for temperature) were removed. Second, Ri and another one with large temperature differences at high
values higher or lower than the mean value ± N standard Ri. Nevertheless, there is no single value of Ri that separates the
deviations computed on a moving window of 500 s were also same populations at the three different heights (note also that
removed. N was set at 7 for wind components and at 8 for the x-axis is logarithmic). In particular, at 32 m, we cannot easily
temperature, values that are good trade-offs for removing most of distinguish different populations of data. Hence, if SBL regimes
the spikes, retaining the valid data. Then a 2D rotation in the mean exist at Dome C, the local Ri seems not to be a good regime
wind coordinates was applied. Data gaps and particularly 12 min predictor.
heating periods were filled in with linearly interpolated values. van Hooijdonk et al. (2015) argue that the behaviour of the
Data were then filtered with a RC-type high-pass filter with a cut- SBL as a whole cannot be driven by a local parameter like z/L or
off frequency of 8 × 10−4 Hz. This filter enabled removal of the Ri (see particularly their figure 9). This point was independently
low-frequency trend within 1 h intervals, particularly the diurnal confirmed by Monahan et al. (2015), who carry out a statistical
cycle. Although a length of time of 8 min sampled a large part diagnosis of the meteorological time series at Cabauw using
of the turbulence spectrum, statistics were not calculated over hidden-Markov model analysis. van Hooijdonk et al. (2015) thus
each 8 min measurement interval, because the effective period suggest trying to characterize the behaviour of the SBL by looking
could be significantly reduced when a large amount of peaks was at external forcings. If the amount of radiative energy that reaches
removed. Variances and covariances were therefore calculated on the surface does not change (no change in cloud cover, for
a 60 min period corresponding to 24 min (3 × 8 min) of effective instance), the SBL is indeed only governed by wind shear and
measurements. Analyses of cospectra and ogives at Dome C surface heat loss. The local stability in the SBL can thus be seen as
showed that the latter period sampled the whole spectrum of a response of these two forcings, not as a driver of SBL dynamics.
turbulence at the four levels on the tower and is thus a good trade- At Dome C, as the sky is often cloud-free, the SBL is expected to
off between integrating turbulent motions and the variability of be primarily driven by the wind shear. In the present study, we
fluxes (not shown). The concatenation of three consecutive 8 min follow Sun et al. (2012), Acevedo et al. (2015) and van Hooijdonk
measurement periods was physically consistent and not affected et al. (2015) and explore the SBL behaviour by looking first at the
by artificial bias at low frequencies, because we first removed wind-speed dependence.
the low-frequency tendency with a high-pass filter. This flux The actual external forcing of the SBL is the wind in
calculation method was verified by carrying out preliminary the free troposphere above the top of the SBL. However,
tests over a snow-covered field in Toulouse, France, during continuous time series of this quantity are hard to obtain from
winter. A good agreement was found between fluxes calculated observations, because they require continuous measurements
from classical measurements and fluxes calculated from data above the boundary-layer height. Here, because we only have
with artificial gaps like 12 min heating periods (not shown access to continuous meteorological measurements in the first
here). tens of metres of the atmosphere at a single location, we wondered
From 1 January 2014–31 October 2015, a Metek USA-1 sonic whether we could find a good proxy of the external wind forcing
thermo-anemometer measured turbulence at a height of 3.5 m in the SBL.
above the ground. The instrument was affected by the cold We could use the wind speed at the highest level of the tower
and failed to function correctly during most of the winter. (41 m). However, in austral summer, i.e. when the boundary
Consequently, only November, December, January and February layer evolves with a clear diurnal cycle, the wind speed in the
(austral summer) data were used in the present study. Technical nocturnal SBL is significantly non-stationary (see Figure 2). It
information on the sonic thermo-anemometer measurements is slowed down near the surface because downward momentum
and data processing are given in Argentini et al. (2005), Pietroni transfer is limited and it accelerates or decelerates aloft due to
et al. (2012) and Vignon et al. (2017). inertial oscillations (Blackadar, 1957; van de Wiel et al., 2010;
Gallée et al., 2015). In particular, the 41 m wind speed increases
2.3. Radiation measurements significantly during the first part of the ‘night’ and is thus not a
good proxy of the wind forcing on the SBL. van de Wiel et al.
Reference radiation measurements at the surface at Dome C are (2012a) showed that, because of the conservation of momentum,
provided by the Baseline Surface Radiation Network (BSRN) there is a characteristic height referred to as the ‘crossing point’, at
(Ohmura et al., 1998; Lanconelli et al., 2011). The surface which the wind speed is relatively constant during the day–night
temperature (Ts ) was retrieved from the measurements of the transition and the following night. This result is a robust feature
downward and upward long-wave fluxes as in Vignon et al. for the nocturnal SBL over flat terrains. Figure 2 shows that,
(2017) using a surface emissivity () of 0.99 (Brun et al., 2012). for Dome C, the crossing point is on average around 10 m.
In this study, all the wind, temperature, turbulence and In austral winter, the SBL is less affected by non-stationarities
radiation data plotted are hourly averages. because of the absence of a diurnal cycle in the polar night.
The whole wind-speed profile in the SBL depends directly on
3. Method for analyzing SBL regimes boundary conditions (external forcings) rather than on the initial
conditions. In conclusion, the 10 m wind speed (hereafter U10m )
As mentioned in the Introduction, evidence for SBL regimes can be considered as the best semi-empirical proxy of the external
in different places of the world can be found in the literature. wind forcing for analysis of the Dome C SBL throughout the
However, the ways of distinguishing and analyzing the regimes whole year.
differ among the various studies. One method is to look at In the present study, we thus first analyze the SBL in relation
internal non-dimensional parameters (e.g. Sorbjan, 1989, 2006) to the 10 m wind speed. Then, we explore the role of a second
and studies like those of Mahrt (1998a) and Grachev et al. (2005) external parameter on the SBL, the incoming radiative energy at
identify regimes using the stability parameter z/L. Mauritsen the surface.


c 2017 Royal Meteorological Society Q. J. R. Meteorol. Soc. 143: 1241–1253 (2017)
1244 E. Vignon et al.

(a) (b) (c)


35 35 35

30 30 30

25 25 25

T17m − Ts (K)

T32m − Ts (K)
T3m − Ts (K)

20 20 20

15 15 15

10 10 10

5 5 5

0 0 0
10–2 100 102 10–2 100 102 10–2 100 102
Ri3m Ri17m Ri32m

Figure 1. Difference in temperature between 3 m (panel a), 17 m (panel b), 32 m (panel c) and the surface versus the local Ri. 2014–2015 data with a stable
stratification are plotted.

3m 10 m 17 m 25 m 32 m 41 m variability of the threshold value itself, i.e. the width of the U10m
range in which the transition occurs, is performed in section 7.1.
T
Situations in which U10m < (>) U10m will be referred to as
3
regime 1 (2), respectively.
In regime 2, the wind speed at each level in Figure 3 correlates
well with the wind speed at 10 m, indicating a full vertical
T
2 connection throughout the tower height. Below U10m , the scatter
increases, indicating that the wind speed at each level becomes
U−U0(m s−1)

significantly less correlated with the value at 10 m. Furthermore,


1 the data align along a less well-defined curve, with a more
upward slope indicating larger vertical wind-speed gradients.
This observation recalls the study of Monahan et al. (2015), in
0 which empirical orthogonal function analysis of Cabauw data
showed two distinct axes along which two different populations
of wind-speed data (two regimes) aligned with the 10 m wind
–1 speed. Paralleling the results of Derbyshire (1999), Banta et al.
(2007) and Acevedo et al. (2015), the two regimes identified in
Figure 3 distinguish a coupled and a decoupled state of the full
–2 SBL with the surface.
0 2 4 6 8 This is further illustrated in Figure 4 at three different heights,
t (h) 3, 17 and 32 m, for three quantities: the ratio of the wind speed
to the near-surface wind speed, the amplitude of the difference
Figure 2. Composite average of the difference between the wind speed and wind
in wind direction with the near-surface wind and the difference
speed at t = 0 h (U0 ) at the six levels of the tower, covering all days in December
and January from January 2011–December 2015. t = 0 h refers to the time at between the temperature and the surface temperature. As the
which the net radiative cooling at the surface starts to be negative. Error bars near-surface wind, we use the wind at 2 m instead of the wind
correpond to ± one standard deviation. The crossing point where the wind is most at 1 m, because no wind measurements were made at 1 m before
constant is around 10 m above the ground. Similar results obtained at Cabauw T
January 2015. In panels a, b and c, for U10m > U10m , the wind-
are presented in van deWiel et al. (2012a).
speed ratios range between 1 and 2 and decrease very slightly
with U10m . As U10m increases, the ratios reach limiting values of
4. Characterization of stable boundary-layer regimes using 1.1 for U3m /U2m , 1.3 for U17m /U2m and 1.5 for U32m /U2m . For
T
the 10 m wind speed U10m < U10m , the range of wind-speed ratios is much larger and
they exhibit no clear dependence on U10m .
4.1. Wind and temperature coupling/decoupling A similar pattern can be observed for the amplitude of the
difference in wind direction in panels e, f and g in Figure 4.
Figure 3 shows scatter plots of the wind speed at six levels versus The U10m threshold distinguishes one U10m range in which the
U10m in stable conditions (identified here when the net radiative difference in wind direction is small (in the right part) and one
flux at the surface is negative and the difference in temperature U10m range in which the variability of the difference in wind
between 2 m and the surface is positive). From right to left in direction at a given wind speed is much larger (left part). It is
each panel, a change in the relation of the wind speed at all worth noting that in 98% of the vertical wind profiles in stable
levels with U10m occurs at U10m ≈ 5–7 m s−1 . As this range is conditions in 2014 and 2015, the wind direction turns anti-
clockwise with height, suggesting the role of the Earth’s rotation
relatively narrow, we will now refer to a wind-speed threshold
T ∗ in the wind veering. Panels e, f and g in Figure 4 thus lead us
U10m , keeping in mind that it is an indicative value rather than
to conclude that the vertical rotation rate and hence the relative
a rigid threshold. A further analysis with respect to any possible
effect of the Coriolis force in the momentum budget is much
more important in regime 1 than in regime 2 over the whole

The value of 6.5 m s−1 is used in Figures 3 and 4 to locate the transition, tower. This observation recalls the SBL regime classification
because this is the value at which most of the regime transitions occur. using the Surface Heat Budget of the Arctic Ocean (SHEBA)


c 2017 Royal Meteorological Society Q. J. R. Meteorol. Soc. 143: 1241–1253 (2017)
Stable Boundary Layer Regimes 1245

(a) (b) z0 < 10−4 m. z0 was retrieved from the Metek sonic thermo-
20 20 anemometer measurements at 3.5 m using the Monin–Obukhov
15 15 relation for momentum:
U2m (m s−1)

U3m(m s−1)
10 10     z 
u∗ zs s
5 5 U(zs ) = ln − ψm , (1)
κ z0 L
0 0
0 5 10 15 0 5 10 15
U10m (m s−1) U10m (m s−1) where zs = 3.5 m, κ is the Von Kármán constant and ψm is the
(c) (d) stability function taken from Holtslag and De Bruin (1988), as in
20 20
Regime 1 Regime 2 Vignon et al. (2017) for stable conditions at Dome C.
15 15
U17m (m s−1)

U25m (m s−1)
Figure 5 first shows that, for a given U10m , u∗ is higher for higher
10 10 z0 (the red crosses are located higher in the graph than the black
5 5 crosses), which could have been expected. Moreover, a regime
0 0
transition for the two subsets is associated with a significant and
0 5 10 15 0 5 10 15 clear change in slope and a U10m threshold marks the low bound
U10m (m s−1) U10m (m s−1)
(e) (f) from which the turbulence starts to be significant and to increase
20 20 consistently with the speed of the wind. Interestingly, this U10m
threshold is lower for high z0 (red crosses) than for low z0 (black
U132m (m s−1)

15 15
U41m (m s−1)

10 10
crosses). It is greater than 5 m s−1 for z0 < 10−4 m, while it is
about 4 m s−1 for z0 > 10−3 m. Similar features can be observed
5 5
for the standard deviation of the turbulent vertical velocity σw
0
0 5 10 15
0
0 5 10 15
(which is less subject to likely horizontal signals from large√non-
U10m (m s−1) U10m (m s−1) turbulent eddies (Mahrt, 1998b; Sun et al., 2012)) and for TKE
(not shown here). The dependence of the transition from SBL
Figure 3. Wind speed at six levels versus U10m . 2014–2015 data with a regimes on the surface roughness is in agreement with the results
stable stratification are plotted. The red dashed line on one panel locates
U10m = 6.5 m s−1 and distinguishes the two regimes. obtained in Derbyshire (1999) and van de Wiel et al. (2012b).
Acevedo et al. (2015) distinguished two SBL regimes using a
‘crossover’ wind-speed threshold at which the vertical gradient
campaign dataset in Grachev et al. (2005) (see also Lüpkes et al., of turbulent quantities reverses sign (positive at low wind
2008). This study concludes that, as z and z/L increase, the SBL speed, negative at high wind speed). This is motivation to
changes progressively from a constant turbulent flux layer to a see whether we could observe the same phenomenon in our
‘supercritical’ stable regime in which the turbulence is intermittent data. We consequently investigated the vertical structure of the
and the wind structure is influenced by the Earth’s rotation, even turbulence. Figure 6 shows bin average plots of u∗ at four heights
very close to the surface. In Figure 4, no information is given versus U10m . The dataset used in this figure corresponds to the
concerning the turbulence activity but, contrary to Grachev et al. two months December 2009–January 2010, for which turbulent
(2005), only two regimes can be identified and the transition is measurements with Atec sonic anemometers were available at
distinguishable by a U10m threshold over the whole tower. four levels on the tower. A significant change in slope is observed
Panels g, h and i in Figure 4 also clearly show the SBL regimes for U10m between 5 and 7 m√s−1 at the four levels. However,
in the temperature difference versus U10m . In regime 2, the Figure 6 and a similar plot of TKE (not shown) do not show a
difference in temperature between the three levels and the surface reversal of the vertical gradient. This suggests that such a gradient
is still moderate, with a slight decrease arising from an increase reversal is not a systematic feature of regime 1 in our dataset.
T Smedman (1988), Mahrt and Vickers (2002) and Sun et al.
in U10m . As one moves from strong U10m to U10m , an abrupt
change in the amplitude of the differences in temperature with (2012) observed top-down propagating turbulence generated by
the surface occurs over the whole tower. The transition from significant wind shear below the peak of low-level jets. At Dome
regime 2 to regime 1 is abrupt, the temperature differences are C, low-level jets generated by inertial oscillation can develop
scattered over a very wide range and reach very high values up during summer ‘nights’ (Barral et al., 2014b; Gallée et al., 2015),
to 30 K between 17 m and the surface. The large variability of the suggesting that top-down propagating turbulence events might
near-surface inversion at a given wind speed in regime 1 suggests occur in summer. In any case, if top-down turbulence events
the intervention of another external forcing in governing the occur at Dome C, they are either too infrequent to be statistically
amplitude of this inversion in low wind-speed conditions. This representative in regime 1 or were not sufficiently well sampled
point is discussed in section 5. We will also see in section 6 that by the tower (in cases of SBL shallower than a few metres, for
this variability can be partly explained by the time dependence of instance) to be visible in Figure 6. The analysis of the occurrence
the near-surface inversion in summer. of top-down propagating turbulence events and the detailed
characterization of the vertical structure of the turbulence at
4.2. Turbulence regimes Dome C are beyond the scope of the present study, but will be
addressed in further research.
In this section, we explore turbulence characteristics in the two
regimes. Figure 5 shows a plot of the friction velocity u∗ measured 4.3. Temperature vertical profiles
with the Metek sonic thermo-anemometer at a height of 3.5 m
during the 2014 and 2015 summer months versus U10m . Vignon et The regime transition at all levels along the tower suggests a
al. (2017) showed that the roughness length for momentum (z0 ) at change of the dynamic and thermodynamic processes in the
Dome C varies with snow properties and varies significantly with SBL, which may affect the shape of the temperature and wind
wind direction, because of a preferential orientation of 5–30 cm vertical profiles. The shape of the vertical temperature profile and
high snow-eroded forms (sastrugi). When the wind is parallel particularly the location of strong temperature inversions are of
to the sastrugi, i.e. for wind directions around 200◦ , z0 drops particular importance for astronomers at Dome C, because they
to 10−5 m. When the wind is perpendicular to the sastrugi, z0 coincide with very strong gradients of the optical properties of the
reaches 5 × 10−3 m. The mean value found in Vignon et al. (2017) air. Previous studies identified different shapes of temperature
was z0 = z0m = 5.6 × 10−4 m. Because of the marked variation profile depending on the wind conditions. van Ulden and Holtslag
in the roughness length, we decided to show two restricted (1985) highlighted ‘exponential-shaped’ temperature profiles in
subsets in Figure 5: one for which z0 > 10−3 m and one for which low wind conditions and ‘convex–concave–convex’ temperature


c 2017 Royal Meteorological Society Q. J. R. Meteorol. Soc. 143: 1241–1253 (2017)
1246 E. Vignon et al.

Figure 4. (a)–(c) Ratio of the wind speed at three levels on the tower over the the wind speed at 2 m versus U10m (in m s−1 ). (d)–(f) Norm of the difference in wind
direction (|δ|) between three levels on the tower and 2 m versus U10m . (g)–(i) Difference in temperature between three levels on the tower and the surface versus
U10m . 2014–2015 data with a stable stratification are plotted. The red dashed line locates U10m = 6.5 m s−1 .

0.6 0.4
z0 >10−3 m 7m
0.35
0.5 z0 < 10−4 m 22 m
0.3 30 m
0.4 0.25 37 m
u∗ (m s−1)
u∗ (m s−1)

0.2
0.3
0.15
0.2 0.1

0.05
0.1
0
0 2 4 6 8 10
0
0 2 4 6 8 10 12 U10m (m s−1)
U10m (m s−1)
Figure 6. u∗ at four levels on the tower versus U10m (1 m s−1 bin averages). The
dataset used here came from the Atec sonic anemometer measurements from 15
Figure 5. u∗ at 3.5 m from the Metek sonic thermo-anemometer versus U10m .
December 2009–31 January 2010 restricted to stable conditions. Note that, for
Red crosses show the subset for which z0 > 10−3 m and black crosses show
better readability, the axes are different from those in Figure 5.
the subset for which z0 < 10−4 m. Only 2014 and 2015 summer data (January,
February, November and December) with stable stratification are plotted.

of the vertical temperature profiles at the regime transition. In


profiles (from bottom to top) with two inflexion points in Figure 7, the solid black lines illustrate two typical examples of
moderate wind conditions at Cabauw. Changes in shape curvature temperature and wind profile in stable conditions at Dome C. On
were observed at Dome C by Genthon et al. (2013), but the 1 September 2014, the wind speed was weak along the whole tower
authors did not search for a link to SBL regimes. In this section, and the temperature profile took on a clear ‘exponential shape’.
we investigate whether there is a systematic change in the shape On 18 March 2014, the wind was stronger and the temperature


c 2017 Royal Meteorological Society Q. J. R. Meteorol. Soc. 143: 1241–1253 (2017)
Stable Boundary Layer Regimes 1247

(a) 1/9/1 4 t = 2300 LT (b) 18/3/14 t = 0300 LT of the high inflexion point of the ‘C–C–C’ shaped temperature
40 40
profiles. For U10m greater than 10 m s−1 , the SBL height and
t-8 the capped temperature inversion of the ‘C–C–C’ temperature
t
30 30 3 profile protrude beyond the top of the tower.
t+8
We now assess whether the temperature profile curvature is a
quantity that is systematically distinguishable from one regime to
z (m)

z (m)
20 20
the other. André and Mahrt (1981) characterized the curvature of
2 the temperature profile using ‘the overall scaled curvature of the
10 10
potential temperature profile’ over the whole SBL. As in many
1 cases we were unable to capture the whole SBL with the tower,
0 0
–80 –70 –60 –50 –40 –80 –70 –60 –50 –40 we calculate the scaled curvature γ along the 45 m tower with a
T (°C) T (°C) similar formula to equation (4) in André and Mahrt (1981):
(c) (d)
40 40 T32m − 2T17m + Ts
γ = , (2)
T32m − Ts
30 30 3

with T32m and T17m the 32 m and 17 m temperature. Negative


z (m)

z (m)

20 20 (positive) γ indicatse convex (concave) temperature profile.


Another way to estimate a local curvature in the temperature
2
10 10 profile is to calculate the local Deacon number Dh (Deacon, 1953;
Lettau, 1979). Here, we calculate it at 13.5 m with the formula
1
0 0  2
0 5 10 15 0 5 10 15
−z ∂∂zT2
U (m s−1) U (m s−1) Dh 13.5m = ∂T
. (3)
Figure 7. (a,b) Vertical temperature profiles and (c,d) vertical wind-speed profiles ∂z 13.5m
along the meteorological tower at two different dates (dates are written as
day/month/year local time). The red (green) line shows the profile 8 h after (before) Details on how this number is calculated with the measurements
the indicated time. The horizontal blue lines and blue numbers distinguish the are given in the Appendix A. Positive (negative) values of Dh
three layers described in Garratt and Brost (1981) for the profiles at time t. correspond to a convex (concave) temperature profile. For a
purely logarithmic profile,
profile was ‘convex–concave–convex’ (hereafter ‘C–C–C’) with
∂ 2T ∂T
two inflexion points, one around 3 m and one around 21 m. −z 2
= , (4a)
Brunt (1934) showed that infrared radiative cooling behaves like ∂z ∂z
a (linear) diffusive process, which gives the temperature profile such that Dh = 1. Basically, γ and Dh are related qualitatively
an exponential shape (see also Cerni and Parish (1984)). The low by sign switch. Figures 8b and 8c show that the two curvature
wind conditions on 1 September 2014 and the exponential shape parameters do not exhibit a clear regime change at a U10m
of the temperature profile thus suggest very weak turbulence threshold.
activity and a radiative-dominated SBL. On the other hand, using Two hypotheses can explain this observation. First, in high
a one-dimensional atmospheric model, Garratt and Brost (1981) wind-speed conditions, the SBL height can exceed the height of
characterized the ‘C–C–C’ temperature profile from the ground the tower and γ underestimates the mean curvature of the whole
to the turbulent SBL height h as a three-layer profile: temperature profile. This may explain why, for U10m greater
(1) A first layer (layer 1, up to 0.1h) with a largely convex than 8 m s−1 , γ diminishes with U10m and Dh 13.5m approaches 1
temperature profile. This layer is dominated by radiative (the constant-flux layer with a logarithmic profile extending to
cooling, because of the difference in emissivity between the 13.5 m). A second hypothesis that could explain the absence
air and the surface. Also, close to the surface, the vertical of a sharp regime transition on the curvature plots is the
temperature gradient is expected to decrease with height, time dependence of the development of the ‘exponential’ and
due to the increasing size of the dominant turbulent eddies. ‘C–C–C’ shapes. Indeed, Estournel and Guedalia (1985) showed
(2) A second layer (0.1h–0.8h) capped by an increasing that under moderate geostrophic wind conditions (10 m s−1 ),
inversion at the highest inflexion point. Because linear the temperature profiles take several hours to develop a ‘C–C–C’
diffusion (such as radiation) tends to smooth gradients, shape. Likewise, the curvature of the ‘exponential’ shape also takes
this layer must be dominated by nonlinear diffusion several hours to develop (Cerni and Parish, 1984). The progressive
by turbulence (Estournel and Guedalia, 1985). This is building of ‘C–C–C’ and ‘exponential’ profiles is clearly visible
essentially in line with Derbyshire (1990)’s theroretical looking at the three curves at time t − 8 h t and t + 8 h in the
considerations on a tendency to ‘inversion singularity’ at upper panels of Figure 7. We also see that the development of the
the top of the SBL. ‘C–C–C’ shape profile is associated with a rise of the inflexion
(3) An uppermost layer (layer 3, 0.8h–h) characterized by a point, sometimes beyond the top of the the tower.
To summarize, probably due to its transient character, the
convex temperature profile, where radiative cooling again
temperature profile curvature does not show a clear separation
dominates.
between two regimes when it is plotted versus the 10 m wind
The three layers are depicted in the right panels in Figure 7. speed. To relate the profile curvature to SBL regimes thus requires
Extending the analysis to all the SBL temperature profiles further investigation, looking for instance at local mixing or time
in 2014–2015, we observed that the location of the largest dependence in complement to external forcings.
curvature in the temperature profile depends to a significant
extent on U10m . Figure 8a shows the 2014–2015 bin averages of 5. Role of the surface radiative energy supply
the normalized local temperature gradients along the tower versus
U10m . For U10m < 5 m s−1 , the normalized temperature gradient In the lower panels of Figure 4, the range of variability of the
decreases steadily with height, reflecting the predominance of temperature and wind vertical gradients in regime 1 is large at a
the ‘exponential’ shape for low wind speeds. For U10m values given wind speed. This suggests that other external forcings play a
between 6.5 and 10 m s−1 , the largest temperature gradients are role in governing the SBL, particularly at low wind speeds. Beyond
found in the upper part of the tower and indicate the presence the wind-speed dependence, the SBL structure, and particularly


c 2017 Royal Meteorological Society Q. J. R. Meteorol. Soc. 143: 1241–1253 (2017)
1248 E. Vignon et al.

(a) the coastal regions of Antarctica over the Plateau (Gallée and
1.8 Gorodetskaya, 2010; Genthon et al., 2013; Barral et al., 2014b;
3m–10m
1.6
Pietroni et al., 2014). Here, we quantify the amount of radiative
10m–17m
energy that comes to the surface by the quantity R+ , defined as
1.4 17m–25m
25m–32m
R+ = |SWdown | − |SWup | +  |LWdown |, (5)
∇zTlocal/∇zTtower

1.2 32m–41m
1 where SWdown and SWup are the downward and upward short-
wave radiative fluxes and LWdown is the downward long-wave
0.8
radiative flux. We use here the quantity R+ rather than the net
0.6 radiative flux Qn , because the latter depends on the surface
temperature. As Qn shares a variable with the near-surface
0.4
inversion, it is thus not an actual external forcing of the system
0.2 composed of the flow and the interactive surface. The lower panels
in Figure 9 show the temperature difference between 10 m and
0
0 2 4 6 8 10 12 the surface versus U10m in stable conditions for 4 subsets of R+ :
U10m (m s−1)
• one in which R+ > 160 W m−2 , corresponding to summer
(b) 1 ‘day–night’ transitions and cloudy summer ‘nights’;
• one in which 120 < R+ < 160 W m−2 , corresponding to
clear-sky summer nights and ‘warming events’ in winter;
0.5 • one in which 80 < R+ < 120 W m−2 , corresponding to
summer–winter transitions and winter days with a
moderate amount of LWdown ; and
• one in which R+ < 80 W m−2 , i.e. days with a very clear
0
γ

sky during the winter polar night.


Figures 9a and 9b show that, for R+ > 120 W m−2 , the tem-
–0.5 perature inversion remains confined to values lower than 15 K.
There are two possible explanations for this observation. First,
the subsets in these two panels include data during the summer
‘day–night’ transitions, i.e data that are not stationary in time and
–1
0 5 10 15 for which inversions are in a developmental or destruction state;
U10m (m s−1) this point is discussed in the following section. Second, in these
conditions, the amount of incoming energy is so large that the
(c)
15 net surface radiative cooling is low and surface energy balance is
reached at a relatively high surface temperature and low temper-
ature inversions even at low wind speed. Figures 9c and 9d show
10 that, for cases with R+ < 120 W m−2 , T10m − Ts can be much
larger at low wind speed, reaching values up to 29 K. The amount
of radiative energy at the surface thus appears to affect the ampli-
tude of the near-surface temperature, particularly in the very stable
Dh 13.5m

5
regime. Figures 9c and 9d also show that the variability of T10m −
Ts at a given wind speed remains large for winds close to the wind-
speed threshold. Moreover, data in Figure 9d, i.e. data correspond-
0
ing to clear-sky conditions during the polar night, look organized
along a ‘reversed’ S curve. This aspect is discussed in section 7.2.
–5
6. Summer versus winter SBL
0 5 10 15
U10m (m s−1) The way we distinguished regimes so far is based on diagnostic
plots as a function of external forcings (or a proxy). Although
Figure 8. (a) Local temperature gradient at each interval on the tower normalized two regimes are diagnostically identifiable, we did not consider
by the temperature gradient over the whole tower versus U10m (0.5 m s−1 bin the time dimension, i.e. how the different quantities, particularly
averages of 2014–2015 data in stable conditions). (b) and (c) γ and Dh 13.5m
versus U10m for stable conditions in 2014–2015. The red line is the 1 line, the
the near-surface temperature inversion, evolve with time in each
green line is the 0 line. regime and during regime transitions. In fact, at Dome C this issue
is closely related to the season we consider. Besides the amount
of radiative energy at the surface, summer and winter SBL also
the near-surface temperature inversion, is expected to depend differ in their temporality. Figure 10 shows the time series of the
on the amount of radiative energy that reaches the surface. temperature difference between 10 m and the surface as well as
More radiative energy toward the surface will cause the surface U10m and R+ during two 33 day periods: one in the 2014–2015
to warm (and the inversion to weaken), or at least reduce its summer (Figure 10a) and one during the polar night in 2015
cooling rate. At Dome C in summer, the Sun remains above (Figure 10b). These periods were specifically chosen because they
the horizon all day long and heats the surface even when the are not affected by major long ‘warming events’.
boundary layer is stable. Therefore, we expect a difference in the On one hand, in summer the ABL is subject to clear diurnal
behaviour of the temperature inversion between the polar night cycles, as visible with the time series of R+ and T10m − Ts
and the summer season. Moreover, although the sky is often in Figure 10a. The daytime convection ‘resets’ the nocturnal
clear, the temperature inversion at Dome C has been shown to stratification between one ‘night’ and the following and the
decrease with the downward long-wave radiation associated with nocturnal SBL enters a regime determined by the conditions at the
the presence of clouds and warm moist air, particularly during ‘day–night’ transition (e.g. van de Wiel et al., 2012a)). Figure 10c
sudden ‘warming events’ i.e. when air masses are advected from shows a scatter plot of T10m − Ts corresponding to the whole


c 2017 Royal Meteorological Society Q. J. R. Meteorol. Soc. 143: 1241–1253 (2017)
Stable Boundary Layer Regimes 1249

(a) (b) regime classification of van Hooijdonk et al. (2015), indicating


30 30
R+ >160W m−2 160 < R+ < 120 W m−2 that the MWST model is quite robust to explain the observations
25 25 T
of the SBL at Cabauw. In our study, U10m basically corresponds
T10m − Ts (K)

T10m − Ts (K)
20 20
to the U10m value below which the ratio |Qn /H| starts to increase
15 15
significantly (see Figure 11a), meaning that, below this value, net
10 10
surface radiative cooling is no longer compensated by H.
5 5
It is thus tempting to assess whether the predictions of MWST
0
0 5 10
0
0 5 10 theory are in agreement with our observations. van de Wiel et al.
U10m (m s−1) U10m (m s−1) (2012a) and van Hooijdonk et al. (2015) derived an expression
(c) (d) for the MWST. First, they considered surfaces with a low heat
30 30
80 < R+ < 120 W m−2 R+ < 80 W m−2 capacity, like snow-covered surfaces, and a dry atmosphere, such
25 25
that the storage term and the latent heat flux can be disregarded
T10m − Ts (K)

T10m − Ts (K)
20 20
in the surface energy budget. Both hypotheses are valid at the first
15 15
order for Dome C conditions. Then, the surface sensible heat flux
10 10
is calculated following the formulation in England and McNider
5 5
(1995) (see also King and Connolley, 1997):
0 0
0 5 10 0 5 10 ⎧
U10m (m s−1) U10m (m s−1) ⎨  −ρc κ2
 p   Uθ (1 − αRb )2 , Rb ≤ 1/α (5a)
H= ln z ln z
z0 z0t
Figure 9. Scatter plots of the temperature difference between 10 m and the surface ⎩
versus U10m for four ranges of R+ . Only 2014–2015 data with a stable stratification
0, Rb > 1/α (5b)
are plotted.

where ρ is the air density, g the gravity acceleration, θ the


period shown in Figure 10a and colours show the local time (LT). difference in potential temperature between z and the surface, cp
The regime transition is clearly visible and one can notice that the heat capacity per unit mass of air and Rb the bulk Richardson
the amplitude of the near-surface temperature inversion during number between the height z and the surface. Here the formula
‘night-time’ (between 1800 and 0700 LT) depends on local time. is slightly modified to account for the difference between the
It is also worth noting that a similar time dependence can be roughness length for momentum z0 and the roughness length for
observed for the vertical wind-speed gradient (not shown here). heat z0t . The maximum sensible heat flux remains greater than or
Maximum near-surface inversion values are reached between equal to the surface energy demand if the wind speed is greater
2200 LT and 0200 LT (dark blue circles in Figure 10c), while than the following threshold:
evening or morning near-surface inversions (cyan circles) are
 1/3
weaker. We also see that the diurnal variations of the near-surface 27 αg |Qn | − |G|
inversion are much larger in the weak-wind regime. A further Umin = z ln(z/z 0 ) ln(z/z 0t ) , (6)
4 θ0 κ 2 ρcp
inspection of the vertical profiles of temperature shows that the
increase of T10m − Ts during the first part of the night is not due with θ0 a reference potential temperature. To calculate Umin at
to a descent of the capping inversion at low heights but is driven Dome C using Eq. (6) can be problematic for three reasons.
by the decrease of surface temperature. First, although snow is a good insulator, the soil heat flux is
On the other hand, during the polar night, the boundary layer not negligible, particularly in summer, but it is not continuously
is almost permanently stably stratified. Except during sudden measured. According to King et al. (2006), the soil heat flux G can
‘warming events’, the boundary layer does not undergo signif- reach one quarter of Qn (in magnitude) in December and January.
icant ‘resetting’ and the SBL progressively transits from one Moreover, Eq. (6) is based on the Monin–Obukhov theory, which
regime to another, depending on the mechanical and radiative is not valid at 10 m in very stable conditions (Vignon et al., 2017).
external forcings. The temperature inversion curve (black line Then, van de Wiel et al. (2012b) point out that the value of Umin
in Figure 10b) clearly shows regime shifts. These regime shifts is sensitive to the surface roughness, which is in agreement with
are mainly associated with variations in U10m (red curve), except the results of section 4.2. The roughness lengths vary significantly
for three transitions (indicated by purple arrows) associated with at Dome C and accurate estimation of Umin requires a prior
peaks of R+ . Strong near-surface inversions correspond to low parametrization of z0 and z0t .
wind speeds and vice versa. These transitions are illustrated further We attempt a first calculation using Eq. (6) for the 2014–2015
in the scatter plot in Figure 10d, in which hourly data of T10m − Ts distribution of Qn in stable conditions and considering values
corresponding to the period shown in Figure 10b are plotted ver- of |G| ranging from 0 W m−2 to 1/4 Qn , different but realistic
sus U10m . Colours indicate the amplitude of the time derivative values of z0 at Dome C, z0t = z0 /10 (Vignon et al., 2017) and
of T10m − Ts in K h−1 . In this figure, data are organized along a α = 4, which is a value that leads to a reasonable fit between
reversed ‘S’ shape like that already observed in Figure 9d. We point measured and estimated sensible heat flux (not shown here).
out that the lower and upper horizontal branch of the reversed This leads to the 10 m Umin distributions in Figure 11b for the
‘S’ correspond to stationary inversions (cyan circles). The vertical period 2014–2015. The black distribution calculated using the
branch of the ‘S’ mainly contains data with non-stationary inver- mean z0 found in Vignon et al. (2017), i.e. z0m = 5.6 × 10−4 m
sions (magenta circles), i.e. data collected during regime shifts. and |G| = 0, is probably the most statistically representative for
Dome C. It peaks at approximately 6 m s−1 , which is in the range
7. Discussion of values of U10m T
identified in section 4. The distribution is
quite sharp (kurtosis equal to 7.0) and the standard deviation is
7.1. Prediction of the minimum wind speed for sustainable equal to 1.1 m s−1 , meaning that the MWST predicts most of the
turbulence (MWST) theory transition to occur not at a fixed U10m value but at wind speeds
close to 6 m s−1 . This is in agreement with our observations. The
van Hooijdonk et al. (2015) clearly demonstrated the existence of full interval of the Umin distribution becomes broader when we
two SBL regimes for wind speeds above or below the MWST, i.e. account for sensitivity to the snow heat flux G and sensitivity to
the minimum wind speed for which the surface sensible heat flux the roughness length. Considering a lower bound for the surface-
H can compensate for the surface energy demand |Qn | − |G|, Qn energy demand term by taking |G| = 14 Qn slightly shifts the
being the net radiative cooling and G the soil heat flux. The results distribution toward lower values (red distribution). To account
obtained in Monahan et al. (2015) independently corroborate the for the variability of the roughness length due to changes in


c 2017 Royal Meteorological Society Q. J. R. Meteorol. Soc. 143: 1241–1253 (2017)
1250 E. Vignon et al.

(a)
R+ (×10−1, W m−2) Summer
35
T10m − Ts (K)
30 U10m (m s−1)

25

20

15

10

–5
December, 1st December, 18th January, 4th
day

(b) Winter
35

30

25

20

15

10

–5
July, 11th July, 28th August, 14th
day

(c) Summer (d)


30 Winter
30 3

25 20:00 LT
25 2.5
20
|dΔT / dt| (K h−1)
20
T10m −Ts (K)

T10m −Ts (K)

15:00 LT 2
15 15
1.5
10 10:00 LT 10
1
5 5
05:00 LT
0 0 0.5

–5 00:00 LT –5 0
0 5 10 0 5 10
U10m (m s−1) U10m (m s−1)

Figure 10. (a,b) Time series of the temperature inversion between 10 m and the surface, U10m , and R+ , during 33 day periods in (a) the 2014–2015 summer and (b)
the 2015 polar night. The green line indicates the 6.5 m s−1 wind-speed value. Purple arrows indicate cloudy conditions (significant peak in LWdown ). Note that the
vertical axis is the same for the three quantities, but the units differ. (c,d) Scatter plots of T10m − Ts versus U10m , corresponding respectively to the periods shown in
(a) and (b). In (c), colours indicate the local time. In (d), colours indicate the amplitude of the time derivative of T10m − Ts (T).

wind direction and snow properties, we show the distributions considering the deviations from Monin–Obukhov similarity
calculated with z0 10 times higher or lower than z0m (green and theory), the Umin distribution can explain why most of the regime
blue distributions, respectively). The peaks of the green, black and transition occurs around 6 m s−1 , but also why some transitions
blue distributions roughly cover a range of 2 m s−1 , illustrating may occur at very different values. This result and Figure 11a
the sensitivity of Umin to roughness length, as already emphasized suggest that the transition between SBL regimes is related to the
in the observations in section 4.2. compensation or non-compensation of the net surface energy
Although the current expression for Umin requires further demand by the sensible heat flux. This allows us to feel confident
improvements to be correctly usable at Dome C (particularly that the fundamental principles of MWST theory can explain


c 2017 Royal Meteorological Society Q. J. R. Meteorol. Soc. 143: 1241–1253 (2017)
Stable Boundary Layer Regimes 1251

(a)
25

20
|Qn/H3.5m|

15

10

0
0 2 4 6 8 10 12 14
U10m (m s−1)
(b)
35 Figure 12. Sketch of the possible equilibrium state of the SBL at Dome C (inspired
z0 = z0m/10, |G| = 0
by Scheffer et al., 2009). Solid (dashed) lines show stable (unstable) equilibria.
30 z0 = z0m × 10, |G| = 0 Dots indicate transition points and arrows show how the SBL behaves if it is not
z0 = z0m, |G| = 14 Qn in equilibrium.
25 z0 = z0m, |G| = 0
in our data prevents us from drawing a definitive conclusion in
20 the present study, Figure 10d opens new perspectives for the
%

conceptual understanding of the SBL. For instance, the likely


15 hysteresis questions the existence of multiple solutions for given
external forcings, i.e. multiple values of T10m − Ts for a given
10 U10m and a given radiative energy supply. This is of particular
importance for predicting the behaviour of the SBL (McNider et
5 al., 1995; van de Wiel et al., 2002). Another study is currently
under way as part of the follow-up theoretical work by the authors
0 on the regime transition in near-surface temperature inversions.
0 2 4 6 8 10 12 14
Umin 10m (m s−1)
7.3. Numerical modelling of the strong near-surface inversions at
Figure 11. (a) Norm of the ratio of the net radiative cooling Qn over the sensible Dome C
heat flux H given by the sonic thermo-anemometer at 3.5 m versus U10m . (b)
2014–2015 frequency distribution of Umin at 10 m calculated with Eq. (6) with The existence of two distinct regimes and the sharp transition
different values of z0 and G. It will be recalled that z0m is the mean value of the between them are three features of the Dome C SBL that
roughness length for momentum found in Vignon et al. (2017) and is equal to
5.6 × 10−4 m. are relevant challenges for the current atmospheric numerical
models. In particular, to reproduce the very strong near-surface
inversions in the low wind regime is an actual litmus test for
the overall existence and transition between the two regimes at physical parametrizations. Beyond the critical role of turbulence,
Dome C. modelling of the very stable regime rests on the representation of
non-turbulent processes like the radiative coupling between the
7.2. Reversed ‘S’ shape dependence of the near-surface temperature atmosphere and the surface. Accurate modelling of the infrared
inversion to the wind speed transfer over the Antarctic Plateau is in itself a challenge,
because it requires a good representation of the moisture
The clear reversed ‘S’ shape observed in Figures 10d and 9d is content in the dry and cold boundary layer, accounting for
intriguing. To our knowledge, this is the first time that evidence for frequent supersaturations with respect to ice (Genthon et al.,
such behaviour of the SBL has been seen in in situ observations. 2017). Moreover, the Antarctic Plateau atmosphere undergoes
In addition, for this particular shape, the ‘S’ seems to ‘fold a continental-scale subsidence associated with the divergence
back’ (also visible in Figure 9d) and, as observed in section 6, of katabatic winds (James, 1989). This subsidence leads to
the vertical branch mainly contains data during regime shifts. an adiabatic heating of the SBL, which can be expected to
These observations could suggest that the SBL may evolve like contribute to the formation of strong temperature inversions
a two-regime dynamical system with a critical transition and (S. De Roode, personal communication). However, the extent
associated with a hysteresis (see Figure 12), as defined in the of such a phenomenon in the energy budget of the Antarctic
review of Scheffer et al. (2009). This kind of critical transition SBL compared with, for instance, the advection of warmer air
recalls the original study by McNider et al. (1995), which evidences from coastal regions, is still an open question. Furthermore, the
transitions between coupled and decoupled states of the SBL using role of likely remnants of dynamical mixing in the very stable
a stability analysis of a two-layer model. Bifurcations between SBL regime, like intermittent turbulence, still needs investigation.
regimes were also studied analytically in van de Wiel et al. (2007) Zilitinkevich and Calanca (2000) and Zilitinkevich (2002) also
and Holdsworth et al. (2016) and, more recently, Donda et al. showed that near-surface mixing in a long-lived SBL (Zilitinkevich
(2015) and van Hooijdonk et al. (2017) found evidence of a critical and Esau, 2005), like the winter SBL at Dome C, can be affected
transition between a turbulent SBL flow and a laminar SBL flow in by non-local features, e.g. the remote influence of downward-
direct numerical simulations. The authors concluded that the SBL propagating internal waves from the free atmosphere. However,
may share similarities with generic dynamical systems like those as no reliable turbulence measurements are available during the
described in Scheffer et al. (2009). In any case, although the scatter polar night, parametrizations of surface turbulent fluxes that


c 2017 Royal Meteorological Society Q. J. R. Meteorol. Soc. 143: 1241–1253 (2017)
1252 E. Vignon et al.

account for such effects are hard to evaluate using current in situ agencies through ‘program CALVA’ (1013) and ‘CoMPASs
observations. project’ is also acknowledged. Further information on the
CALVA meteorological programme is available at the website
8. Conclusions http://lgge.osug.fr/∼genthon/calva/home.shtml.

The analysis of a meteorological dataset along a 45 m tower, Appendix A: Deacon number for temperature
complete with radiation and turbulence measurements, revealed
the existence of two SBL regimes at Dome C on the Antarctic The Deacon number for temperature at 13.5 m is calculated using
Plateau. Because the behaviour of the SBL is driven primarily a discrete form of Eq. (3) considering the available measurements
by wind shear, the regime transition can be characterized with a on the tower. Here, the two terms ∂T/∂z and ∂ 2 T/∂z2 of Eq. (3)
10 m wind-speed threshold. For winds above the threshold, the are developed:
wind is vertically coupled over the 45 m tower and probably over
the whole SBL. Temperature inversions are moderate and the ∂T T17m − T10m
= , (A1)
turbulence activity is significant and increases steadily with an ∂z 13.5m z17m − z10m
increase in wind strength. For 10 m winds below the threshold,
the wind and temperature gradients increase dramatically and the
continuous turbulence cuts off. In these conditions, temperature ∂ 2T
differences between 10 m and the surface can reach values greater ∂z2 13.5m
 
than 25 K. Weak-wind conditions are generally associated with T25m − T10m T17m − T3m
‘exponential’ shaped vertical temperature profiles and moderate = − /(z17m − z10m ). (A2)
z25m − z10m z17m − z3m
wind conditions with ‘C–C–C’ shaped vertical temperature
profiles. However, because vertical temperature profiles are References
usually non-stationary in time, direct use of their curvature as a
regime indicator is not recommended. Paralleling results obtained Acevedo OC, Mahrt L, Puhales FS, Costa FD, Medeiros LE, Degrazia GA.
at Cabauw, the regime transition appears to be closely related to 2015. Contrasting structures between the decoupled and coupled states
of the stable boundary layer. Q. J. R. Meteorol. Soc. 142: 693–702.
the balance or imbalance between the surface sensible heat flux https://doi.org/10.1002/qj.2693.
and net surface cooling. This observation and the values of the André JC, Mahrt L. 1981. The nocturnal surface inversion and influence of
wind-speed thresholds are in good agreement with predictions by clear air radiative cooling. J. Atmos. Sci. 39: 864–878.
the MWST theory. The amplitude of the near-surface temperature Argentini S, Viola A, Sempreviva AM, Petenko I. 2005. Summer boundary-layer
inversion was also shown to depend on the amount of radiative height at the plateau site of Dome C, Antarctica. Boundary Layer Meteorol.
115: 409–422. https://doi.org/10.1007/s10546-004-5643-6.
energy that reaches the surface. Moreover, the structure of the Aristidi E, Agabi K, Azouit M, Fossat E, Vernin J, Travouillon T, Lawrence JS,
summertime SBL is shown to depend on local time, particularly Meyer C, Storey JWV, Halter B, Roth WL, Walden V. 2005. An analysis of
in the low wind-speed regime. temperatures and wind speeds above Dome C, Antarctica. Astron. Astrophys.
The study of the regime transition in the clear-sky polar night 430: 739–746. https://doi.org/10.1051/0004-6361:20041876.
Banta RM, Mahrt L, Vickers D, Sun J, Balsey BB, Pichugina YL. 2007. The very
and observation of the ‘S’ shape dependence of the near-surface stable boundary layer on nights with weak low-level jets. J. Atmos. Sci. 64:
temperature on the 10 m wind speed suggest that the SBL at 3068–3090.
Dome C obeys a two-regime system with a critical transition. This Barral H, Genthon C, Trouvilliez A, Brun C, Amory C. 2014a. Blowing snow
observation provides new insights for the conceptual modelling at D17, Adélie Land, Antarctica: Atmospheric moisture issues. Cryosphere
of SBL dynamics. To reproduce the two regimes and very 8: 1905–1919. https://doi.org/10.5194/tc-8-1905-2014.
Barral H, Vignon E, Bazile E, Traullé O, Gallée H, Genthon C, Brun C,
large temperature inversions is also a challenge for atmospheric Couvreux F, Le Moigne P. 2014b. ‘Summer diurnal cycle at Dome C on the
numerical models, given that this certainly requires not only an Antarctic Plateau’. In Proceedings of the AMS 21st Symposium on Boundary
accurate turbulent closure but also a radiative transfer scheme Layers and Turbulence. Leeds, UK.
adapted to the cold dry air over the Antarctic Plateau. Analyses of Blackadar AK. 1957. Boundary layer wind maxima and their significance for
the growth of nocturnal inversions. Bull. Am. Meteorol. Soc. 38: 283–290.
model performance over Dome C are under way in the framework
Brun E, Six D, Picard G, Vionnet V, Arnaud L, Bazile E, Boone A, Bouchard
of the GEWEX Atmospheric Boundary Layers Experiment 4.† A, Genthon C, Guidars V, Le Moigne P, Rabier F, Seity Y. 2012.
Finally, this first climatology of SBL regimes opens the way Snow/atmosphere coupled simulation at Dome C, Antarctica. J. Glaciol.
for further in-depth analysis, considering for instance non- 52(204): 721–736.
dimensionnal analysis of the flow in relation with turbulence Brunt D. 1934. Physical and Dynamical Meteorology. Cambridge University
Press: Cambridge, UK.
ignition/extinction. The time evolution of the SBL in each regime Casasanta G, Pietroni I, Petenko I, Argentini S. 2014. Observed and modelled
also remains to be investigated, exploring in more detail the convective mixing-layer height at Dome C, Antarctica. Boundary Layer
evening transitions in summer, but also sudden ‘warming events’, Meteorol. 151: 587–608. https://doi.org/10.1007/s10546-014-9907-5.
which act like ‘reset buttons’ for the winter SBL. Cerni TA, Parish TR. 1984. A radiative model of the stable nocturnal boundary
layer with application to the polar night. J. Clim. Appl. Meteorol. 23:
1563–1572.
Acknowledgements Conangla L, Cuxart J, Soler M. 2008. Characterisation of the nocturnal
boundary layer at a site in northern Spain. Boundary Layer Meteorol. 128:
We are grateful to Hubert Gallée, Arnold Moene, Chiel 255–276. https://doi.org/10.1007/s10546-008-9280-3.
Deacon EL. 1953. Vertical Profiles of Mean Wind in the Surface Layer of the
van Heerwaarden and Stephan de Roode for many fruitful Atmosphere, Meteorological Office. Geophysical memoirs No 91: London,
discussions and to three anonymous referees for very helpful UK.
comments. We gratefully acknowledge Christian Lanconelli and Derbyshire SH. 1990. Nieuwstadt’s stable boundary layer revisited. Q. J. R.
Maurizio Busetto for dissemination of the BSRN data and Meteorol. Soc. 126: 127–158.
Guylaine Canut for her help in processing data from the Derbyshire SH. 1999. Boundary-layer decoupling over cold surfaces as a
physical boundary-instability. Boundary Layer Meteorol. 90: 297–325.
sonic thermo-anemometers. We acknowledge financial support Donda JMM, Van Hooijdonk IGS, Moene AF, Van Heijst GJF, Clercx HJH.
provided by Bas van de Wiel’s ERC-Consolidator grant (648666), 2015. The maximum sustainable heat flux in stably stratified channel flows.
which enabled a visit to his group by the first author. This Q. J. R. Meteorol. Soc. 142: 781–792.
research was supported by INSU (programs LEFE CLAPA England DE, McNider RT. 1995. Stability functions based upon shear functions.
Boundary Layer Meteorol. 74: 113–130.
and GABLS4, OSUG (GLACIOCLIM observatory). Logistical
Estournel C, Guedalia D. 1985. Influence of geostrophic wind on atmospheric
support by the French (IPEV) and Italian (PNRA) polar nocturnal cooling. J. Atmos. Sci. 42: 2695–2698.
Gallée H, Gorodetskaya I. 2010. Validation of a limited area model over
Dome C, Antarctic Plateau, during winter. Clim. Dyn. 23: 61–72.

http://www.cnrm-game-meteo.fr/aladin/meshtml/GABLS4/GABLS4.html https://doi.org/10.1007/s00382-008-0499-y.


c 2017 Royal Meteorological Society Q. J. R. Meteorol. Soc. 143: 1241–1253 (2017)
Stable Boundary Layer Regimes 1253

Gallée H, Guyomarc’h G, Brun E. 2001. Impact of snow drift on the Antarctic ice Newsom RK, Banta RM. 2003. Shear-flow instability in the stable nocturnal
sheet surface mass balance: Possible sensitivity to snow surface properties. boundary layer as observed by a Doppler lidar during CASE-99. J. Atmos.
Boundary Layer Meteorol. 99: 1–19. Sci. 60: 16–33.
Gallée H, Barral H, Vignon E, Genthon C. 2015. A case study of Ohmura A, Gilgen H, Hegner H, Müller G, Wild M, Dutton EG, Forgan
a low-level jet during OPALE. Atmos. Chem. Phys. 15: 6237–6246. B, Fröhlich C, Philipona R, Heimo A, König-Langlo G, McArthur B,
https://doi.org/10.5194/acp-15-1-2015. Pinker R, Whitlock CH, Dehne K. 1998. Baseline surface radiation network
Garratt JR, Brost RA. 1981. Radiative cooling effects within and above the (BSRN/WMRC), a new precision radiometry for climate research. Bull. Am.
nocturnal boundary layer. J. Atmos. Sci. 38: 2370–2745. Meteorol. Soc. 79: 2115–2136.
Genthon C, Town MS, Six D, Favier V, Argentini S, Pellegrini A. 2010. Petenko I, Argentini S, Pietroni I, Viola A, Mastrantonio G, Casasanta G,
Meteorological atmospheric boundary layer measurements and ECMWF Aristidi E, Bouchez G, Agabi A, Bondoux E. 2014. Observations of optically
analyses during summer at Dome C, Antarctica. J. Geophys. Res. 115: active turbulence in the planetary boundary layer by sodar at the Concordia
D05104. https://doi.org/10.1029/2009JD012741. astronomical observatory, Dome C, Antarctica. Astron. Astrophys. 568: A44.
Genthon C, Six D, Gallée H, Grigioni P, Pellegrini A. 2013. Two years https://doi.org/10.1051/0004-6361/201323299.
of atmospheric boundary layer observations on a 45-m tower at Dome Pietroni I, Argentini S, Petenko I, Sozzi R. 2012. Measurements
C on the Antarctic Plateau. J. Geophys. Res. Atmos. 118: 3218–3232. and parametrizations of the atmospheric boundary-layer height
https://doi.org/10.1002/jgrd.50128. at Dome C, Antarctica. Boundary Layer Meteorol. 143: 189–206.
https://doi.org/10.1007/s10546-011-9675-4.
Genthon C, Piard L, Vignon E, Madeleine J-B, Casado M, Gallée H. 2017.
Atmospheric moisture supersaturation in the near-surface atmosphere Pietroni I, Argentini S, Petenko I. 2014. One year of surface-based temperature
at Dome C, Antarctic Plateau. Atmos. Chem. Phys. 17: 691–704. inversions at Dome C, Antarctica. Boundary Layer Meteorol. 150: 131–151.
https://doi.org/10.5194/acp-17-691-2017. https://doi.org/10.1007/s10546-013-9861-7.
Grachev AA, Fairall CW, Persson P, Andreas EL, Guest PS. 2005. Stable Ricaud P, Genthon C, Durand P, Attié J, Carminati F, Canut G, Vanacker J,
boundary-layer scaling regimes: The SHEBA data. Boundary Layer Meteorol. Moggio L, Courcoux Y, Pellegrini A, Rose T. 2012. Summer to winter diurnal
116: 201–235. https://doi.org/10.1007/s10546-004-2729-0. variabilities of temperature and water vapour in the lowermost troposphere
as observed by HAMSTRAD over Dome C, Antarctica. Boundary Layer
Holdsworth AM, Rees T, Monahan AH. 2016. Parameterization sensitivity Meteorol. 143: 227–259.
and instability characteristics of the maximum sustainable heat flux
Scheffer M, Bascompte J, Brock W, Brovkin V, Carpenter S, Dakos V, Held H,
framework for predicting turbulent collapse. J. Atmos. Sci. 73: 3527–3540.
van Nes E, Rietkerk M, Sugohara G. 2009. Early-warming signals for critical
https://doi.org/10.1175/JAS-D-16-0057.1.
transitions. Nature 461: 591–596. https://doi.org/10.1038/35098000.
Holtslag AAM, De Bruin HAR. 1988. Applied modeling of the nighttime
Smedman AS. 1988. Observations of a multi-level turbulence structure in a very
surface energy balance over land. J. Appl. Meteorol. 27: 689–703.
stable atmospheric boundary layer. Boundary Layer Meteorol. 44: 231–253.
Holtslag A, Svensson G, Baas P, Basu S, Beare B, Beljaars A, Bosveld F, Sorbjan Z. 1989. Structure of the Atmospheric Boundary Layer. Prentice-Hall:
Cuxart J, Lindvall J, Steeneveld G, Tjernström M, Van De Wiel B. 2013. New Jersey.
Stable Atmospheric Boundary Layers and Diurnal Cycles: Challenges for
Sorbjan Z. 2006. Local structure of turbulence in stably stratified boundary
Weather and Climate Models. Bull. Amer. Meteor. Soc. 94: 1691–1706.
layers. J. Atmos. Sci. 63: 1526–1537. https://doi.org/10.1175/JAS3704.1.
https://doi.org/10.1175/BAMS-D-11-00187.1.
Sun J, Mahrt L, Banta RM, Pichugina YL. 2012. Turbulence regimes and
van Hooijdonk IGS, Donda J, Clercx H, Bosveld F, van de Wiel BJH. 2015.
turbulence intermittency in the stable boundary layer during CASES-99. J.
Shear capacity as prognostic for nocturnal boundary-layer regimes. J. Atmos.
Atmos. Sci. 69: 338–351. https://doi.org/10.1175/JAS-D-11-082.1.
Sci. 72: 1518–1532. https://doi.org/10.1175/JAS-D-14-0140.1.
Sun J, Nappo CJ, Mahrt L, Belušić D, Grisogono B, Stauffer DR, Pulido
van Hooijdonk IGS, Moene AF, Scheffer M, Clercx H, Van de Wiel BJH. 2017. M, Staquet C, Jiang Q, Pouquet A, Yagüe C, Galperin B, Smith RB,
Early Warning Signals for Regime Transition in the Stable Boundary Finnigan JJ, Mayor SD, Svensson G, Grachev AA, Neff WD. 2015. Review
Layer: A Model Study. Boundary-Layer Meteorol. 162(2): 283–306. of wave–turbulence interactions in the stable atmospheric boundary layer.
https://doi.org/10.1007/s10546-016-0199-9. Rev. Geophys. 53: 956–993. https://doi.org/10.1002/2015RG000487.
James IM. 1989. The Antarctic drainage flow: Implications for hemispheric van Ulden AP, Holtslag AAM. 1985. Estimation of atmospheric boundary
flow on the southern hemisphere. Antarct. Sci. 1: 279–290. layer parameters for diffusion applications. J. Clim. Appl. Meteorol. 24:
King JC, Connolley WM. 1997. Validation of the surface energy balance over 1196–1207.
the Antarctic Ice Sheet in the U.K. Meteorological Office unified climate van de Wiel BJH, Moene WH, Hartogensis OK, De Bruin HAR, Holtslag AAM.
model. J. Clim. 10: 1273–1287. 2002. Intermittent turbulence in the stable boundary layer over land. Part
King JC, Argentini SA, Anderson PS. 2006. Contrasts between the II: A system dynamics approach. J. Atmos. Sci. 59: 2567–2581.
summertime surface energy balance and boundary layer structure at van de Wiel BJH, Moene AF, Hartogensis OK, De Bruin HAR, Holtslag AAM.
Dome C and Halley stations, Antarctica. J. Geophys. Res. 111: D02105. 2003. Intermittent turbulence in the stable boundary layer over land. Part
https://doi.org/10.1029/2005JD006130. III: A classification for observations during CASES-99. J. Atmos. Sci. 28:
Lanconelli C, Busetto M, Dutton EG, König-Langlo G, Maturilli M, Sieger 2509–2522.
R, Vitale V, Yamanouchi T. 2011. Polar baseline surface radiation van de Wiel BJH, Moene AF, Steeneveld GJ, Hartogensis OK, Holtslag AAM.
measurements during the international polar year 2007–2009. Earth Syst. 2007. Predicting the collapse of turbulence in stably stratified boundary
Sci. Data Discuss. 3: 1–8. https://doi.org/10.5194/essd-3-1-2011. layers. Flow Turbul. Combust. 79: 251–274. https://doi.org/10.1007/s10494-
Lettau HH. 1979. Wind and temperature profile prediction for diabatic 007-9094-2.
surface layers including strong inversion cases. Boundary Layer Meteorol. van de Wiel BJH, Moene AF, Steeneveld GJ, Baas P, Bosveld FC, Holstlag AAM.
17: 443–464. 2010. A conceptual view on inertial oscillations and nocturnal low-level jets.
Libois Q, Picard G, Arnaud L, Morin S, Brun E. 2014. Modeling J. Atmos. Sci. 67: 2679–2689. https://doi.org/10.1175/2010JAS3289.1.
the impact of snow drift on the decameter-scale variability of snow van de Wiel BJH, Moene AF, Jonker HJJ. 2012a. The cessation of continuous
properties on the Antarctic Plateau. J. Geophys. Res. Atmos. 119: 662–681. turbulence as precursor of the very stable boundary layer. J. Atmos. Sci. 69:
https://doi.org/10.1002/2014JD022361. 3116–3127. https://doi.org/10.1175/JAS-D-12-0107.1.
Lüpkes C, Vihma T, Birnbaum G, Wacker U. 2008. Influence of leads in sea ice van de Wiel BJH, Moene AF, Jonker HJJ, Baas P, Basu S, Donda JMM,
on the temperature of the atmospheric boundary layer during polar night. Sun J, Holtslag AAM. 2012b. The minimum wind speed for sustainable
Geophys. Res. Lett. 35: L03805. https://doi.org/10.1029/2007GL032461. turbulence in the nocturnal boundary layer. J. Atmos. Sci. 69: 3097–3115.
McNider RT, England DE, Friedman MJ, Shi X. 1995. Multiple regimes of https://doi.org/10.1175/JAS-D-12-064.1.
wind, stratification, and turbulence in the stable boundary layer. J. Atmos. Vickers D, Mahrt L. 1997. Quality control and flux sampling problems for
Sci. 52: 1602–1614. https://doi.org/10.1175/1520-0469. tower and aircraft data. J. Atmos. Oceanic Technol. 14: 512–526.
Mahrt L. 1998a. Nocturnal boundary-layer regimes. Boundary Layer Meteorol. Vignon E, Genthon C, Barral H, Amory C, Picard G, Gallée H, Casasanta G,
88: 255–278. https://doi.org/10.1023/A:1001171313493. Argentini S. 2017. Momentum and heat flux parametrization at Dome C,
Mahrt L. 1998b. Stratified atmospheric boundary layers and breakdown of Antarctica: A sensitivity study. Boundary Layer Meteorol. 162(2): 341–367.
models. Theor. Comput. Fluid Dyn. 11: 263–279. https://doi.org/10.1007/s10546-016-0192-3.
Mahrt L. 2014. Stably stratified atmospheric boundary layers. Annu. Rev. Fluid Zilitinkevich S. 2002. Third-order transport due to internal waves and non-
Mech. 46: 23–45. https://doi.org/10.1146/annurev-fluid-010313-14135. local turbulence in the stably stratified surface layer. Q. J. R. Meteorol. Soc.
Mahrt L, Vickers D. 2002. Contrasting vertical structures of nocturnal boundary 128: 913–925.
layers. Boundary Layer Meteorol. 105: 351–363. Zilitinkevich S, Calanca P. 2000. An extended similarity theory for the stably
Mastrantonio G, Malvestuto V, Argentini S, Georgiadis T, Viola A. 1999. stratified atmospheric surface layer. Q. J. R. Meteorol. Soc. 126: 1913–1923.
Evidence of the convective boundary layer developing on the Antarctic Zilitinkevich S, Esau I. 2005. Resistance and heat-transfer laws for stable and
Plateau during the summer. Meteorol. Atmos. Phys. 71: 127–132. neutral planetary boundary layers: Old theory advanced and re-evaluated.
Mauritsen T, Svensson G. 2007. Observations of stably stratified shear-driven Q. J. R. Meteorol. Soc. 131: 1863–1892. https://doi.org/10.1256/qj.04.143.
atmospheric turbulence at low and high Richardson numbers. J. Atmos. Sci. Zilitinkevich S, Elperin T, Kleeorin N, Rogachevskii I, Esau I, Mauritsen
64: 645–655. https://doi.org/10.1175/JAS3856.1. T, Miles MW. 2008. Turbulence energetics in stably stratified geophysical
Monahan AH, Rees T, He Y, McFarlane N. 2015. Multiple regimes of wind, flows: Strong and weak mixing regimes. Q. J. R. Meteorol. Soc. 134: 793–799.
stratification, and turbulence in the stable boundary layer. J. Atmos. Sci. 72: https://doi.org/10.1002/qj.264.
3178–3198. https://doi.org/10.1175/JAS-D-14-0311.1.


c 2017 Royal Meteorological Society Q. J. R. Meteorol. Soc. 143: 1241–1253 (2017)

You might also like