Local and non-local controls on a persistent cold-air pool in the Arve River Valley G. Arduinia,b,∗, C. Chemela,c and C. Staquetb aCentre for Atmospheric & Instrumentation Research, University of Hertfordshire, College Lane, Hatfield, AL10 9AB, UK bUniversite´ Grenoble Alpes, CNRS, LEGI, F-38000, Grenoble, France cNational Centre for Atmospheric Science, Centre for Atmospheric & Instrumentation Research, University of Hertfordshire, College Lane, Hatfield, AL10 9AB, UK ∗Correspondence to: G. Arduini, European Centre for Medium-Range Weather Forecasts, Reading, United Kingdom E-mail: gabriele.arduini@ecmwf.int A numerical model is used to simulate a persistent cold-air pool (PCAP) event that occurred in the section of the Arve River Valley around Passy in the French Alps. During this period, an upper-level ridge from the Atlantic moved over Europe, allowing a PCAP to form and persist over time. The impact of the upper-level ridge on the PCAP and on the dynamics within the valley section is quantified, by examining the mass and heat budgets of the valley atmosphere. During the persistent stage, the magnitude of the flow through the tributary valleys is enhanced by the large-scale flow. Also, the direction of the flow through one of the tributaries is found to be determined by the height of the PCAP with respect to that of the tributary above the valley floor. The tributary flows, together with subsiding motions at the valley top, control by and large the nighttime valley-scale circulation and the thermal structure of the upper part of the PCAP, whereas thermally-driven valley flows control its lower part. When the upper- level ridge passes over the Arve River Valley, warm air advection through the tributaries continuously erodes the upper part of the PCAP during nighttime, thereby reducing its depth, while down-valley flows export the air mass out of the valley. As the ridge moves away from the valley, the near-surface air is found to be trapped within the valley. This trapping results from the advection of warm air in the upper part of the PCAP by the large-scale flow channelled through one of the tributary. This reduces the thermally-induced pressure difference in the down-valley direction, thereby suppressing the near-surface down-valley flow. The study therefore highlights the interplay between the large-scale flow, the tributary flows and the thermal structure of the PCAP. Key Words: PASSY-2015; cold-air pool; complex terrain; stable boundary layer 1. Introduction A cold-air pool (CAP) is a stably stratified layer of air confined at the bottom of a valley or basin. Diurnal CAPs form frequently during nocturnal hours under clear-sky and weak synoptic wind conditions and are destroyed following sunrise mainly by convection resulting from surface heating. When relatively warm air associated with anticyclonic conditions is transported right above deep terrain over multiple days (Reeves and Stensrud 2009), convection is often insufficient to destroy the stable layer within the valley, particularly during wintertime (see for instance Vrhovec and Hrabar 1996; Lu and Zhong 2014), allowing a persistent cold-air pool (PCAP) to develop. Since a PCAP is characterised by a long-lasting (multi-day) valley inversion and weak winds, air pollutants and moisture emitted within the PCAP are trapped and accumulate from day to day, thereby affecting air quality (Silcox et al. 2012; Chemel et al. 2016; Largeron and Staquet 2016a) and visibility when fog forms. The removal of the PCAP may occur because of the passage of a synoptic disturbance over the valley (Zhong et al. 2001), turbulent erosion of the PCAP from the top (Lareau and Horel 2015a) or the displacement of the cold-air layer out of the valley (Flamant et al. 2006; Lareau and Horel 2015b). While the formation, persistence and destruction stages are common to PCAP in many different regions and conditions (Zhong et al. 2001; Za¨ngl 2005a; Lareau et al. 2013; Lu and Zhong 2014; Largeron and Staquet 2016b), the characteristics of the PCAP (depth and strength) during each stage depend on the characteristics of the surrounding terrain and variability of the synoptic conditions. Thereafter a distinction is made between the processes that control the characteristics of the PCAP and are governed by those of the terrain at the valley or basin scale, which are referred to as local controls (such as This article is protected by copyright. All rights reserved. A cc ep te d A rti cl e This article has been accepted for publication and undergone full peer review but has not been through the copyediting, typesetting, pagination and proofreading process which may lead to differences between this version and the Version of Record. Please cite this article as doi: 10.1002/qj.3776 local thermally-driven flows), and processes affecting the PCAP characteristics but occurring at scales larger than that of the valley (such as the synoptic wind above the valley boundary layer and flows from surrounding tributary valleys modified by the large- scale flow), which are referred to as non-local controls. Previous work has considered local controls on CAPs. Vrhovec (1991) showed using results from numerical simulations of a diurnal CAP in the Slovenj Gradec basin that a dynamical decoupling can take place between the flow within the valley boundary layer and the synoptic flow aloft. As a result, at the top of a CAP an abrupt change in wind speed and direction was observed. In such decoupled condition thermally-driven along- slope and along-valley flows are the key dynamical features of the valley atmosphere (Largeron and Staquet 2016a). However, much remains to be understood on their role in the formation and evolution of CAPs (Bodine et al. 2009). In a numerical modelling case study of the formation of a CAP in a ∼ 100-m deep narrow valley, Vosper et al. (2014) concluded that the principal mechanism responsible for the formation of the CAP was the sheltering effect of the valley. The sheltering was shown to increase the turbulent heat flux divergence close to the ground compared to a less sheltered region (i.e. hilltop sites), thereby increasing near-surface cooling and promoting the formation of a ground-based inversion. However, in such a shallow valley, thermally-driven flows are rapidly suppressed after the evening transition (Clements et al. 2003; Vosper et al. 2014). In a deep and narrow valley dynamically decoupled from the atmosphere above, downslope flows can persist for a longer time, reaching a quasi-steady state if a down-valley flow can develop during the night (Burns and Chemel 2015; Arduini et al. 2016). Under these conditions, downslope flows are the main driver of the sensible heat flux divergence, which enhances the cooling of the atmosphere in the valley with respect to a flat region nearby. Downslope flows are also important in the evolution of the thermal structure of the upper part of a CAP, through the vertical transport induced by the air flowing down the slopes (Whiteman 1982; Kiefer and Zhong 2011; Burns and Chemel 2015; Arduini et al. 2016). In addition to thermally-driven flows, the variations of the orography and surface properties affect the evolution of a CAP. Because of the reduction in the volume of atmosphere in a valley or basin when compared to that over a flat terrain, a valley or basin experiences a cooling that is stronger than if it were a flat terrain for the same amount of heat loss (Whiteman 1990). A differential cooling can also take place between different sections of the same valley because of along-valley geometrical variations (McKee and O’Neal 1989). This differential cooling can lead to along-valley flows that are up-valley instead of down-valley during the night (Arduini et al. 2017), affecting the mass and heat budgets of the valley atmosphere. If the ground is covered by snow (as is common in winter in alpine regions), surface albedo is increased and the heat conductivity to the ground is significantly decreased, leading to stronger ground-based inversions (Za¨ngl 2005b; Billings et al. 2006; Neemann et al. 2015). Non-local controls on PCAPs have been examined for a number of case studies. Zhong et al. (2001) showed that the formation and maintenance of a PCAP event in the Columbia basin in the Pacific Northwest region of North America was primarily the result of the continuous advection of warm air immediately above the basin by strong westerly winds descending the lee slopes of the surrounding Cascade mountains. Similar results as regards the importance of large-scale advection of warm air above the terrain in the formation and maintenance of a PCAP in the Salt Lake Valley, Utah, USA, were reported by Wei et al. (2013). Lu and Zhong (2014) quantified the respective contributions of the different terms determining the heat budget of the atmosphere during this PCAP event. Large-scale advection was found key to the formation, maintenance and destruction of the PCAP. While these studies have demonstrated the essential role of warm air advection in the formation and maintenance of a PCAP, the dynamical controls of large-scale flows on the temporal evolution of CAPs have received less attention. In a numerical modelling case study Za¨ngl (2005a) showed the importance of the direction of the synoptic pressure gradient and of the along- valley terrain geometry, with respect to the Alpine foreland, in the maintenance of CAPs in a deep Alpine valley system. Model results indicated that the direction of the synoptic pressure gradient controlled the direction of the along-valley drainage flows. The author concluded that, for that particular case study, the ideal conditions for the removal of a cold-air pool are met when the synoptic pressure gradient is orientated down a constriction in the along-valley direction. In a following work based on semi- idealised numerical simulations and observations, Za¨ngl (2005c) extended this analysis to the shallower Bavarian Danube basin. Model results showed that advection resulting from the synoptic pressure difference (pressure-driven channelling) and advection down the direction of the synoptic wind (forced-channelling) can both act to remove the PCAP, depending on the direction of the synoptic wind with respect to the orientation of the highest mountain range. Another dynamical mechanism that may affect the temporal evolution of PCAPs is shear-induced turbulent mixing at the top of the PCAP due to strong winds aloft. Some studies questioned the effectiveness of such turbulent erosion for the destruction of CAPs over time-scales shorter than one day (see for instance Zhong et al. 2001, 2003). However, semi-idealised numerical simulations analysed by Za¨ngl (2005c) suggested that this mechanism can contribute significantly to the destruction of a CAP in the relatively shallow Danube basin. In deep valleys, idealised simulations by Lareau and Horel (2015a) indicated that CAPs can be removed by turbulent erosion on a time-scale of less than a day, provided that the speed of the flow aloft increases continuously with time such that the Froude number exceeds a threshold value of about 2. A more systematic study on the link between the size of the valley and the synoptic conditions have been performed by Sheridan (2019) by means of idealised 2D simulations. The author showed that for deep ( 1000 m-deep) valleys embedded within a plateau, for increasing background stability, the CAP becomes increasingly more sheltered from the external flow and so more prone to being persistent. This behaviour was found to be well described by a non-dimensional valley depth parameter (Vosper and Brown 2008). However, in the case of a valley embedded between two hills, a more common situation in the Alpine environment, the relationship between the non-dimensional valley depth and CAP persistency was found more complex, and a very strong background stability (in a real case associated for instance with synoptic warm air advection) was required to maintain the CAP persistent in time. The contribution of tributary flows in the mass budget of a valley was studied extensively as part of the Advanced Studies in Complex Terrain (ASCOT) programme (Clements et al. 1989). However, the computation of mass fluxes through tributaries relied solely on limited observations, and so several assumptions and approximations had to be made to estimate their contributions to the mass budget. Porch et al. (1989) calculated that the mass flux through a single tributary may account for 5 to 15% of the drainage out of Brush Creek Valley, Colorado, USA. Assuming a similar contribution from all the tributaries of Brush Creek Valley, the authors estimated that tributary flows control by and large the mass budget of the valley, and so that the contribution of slope flows and subsiding motions at the valley top are only minor. Numerical model simulations reported by O’Steen (2000) for an idealised valley-tributary system under dynamically decoupledThis article is protected by copyright. All rights reserved. A cc ep te d A rti cl e Figure 1. (a) Topographic map of domain d02 with the positions of the nested domains d03, d04 and d05, represented by continuous-line boxes. The dashed-line box represents the region covered by the semi-idealised simulation, WRF-I. (b) Topographic map of the innermost domain d05; the dashed black polygon indicates the horizontal extent of the PASSY control volume, the white line indicates the path of cross-sections used in the paper; mass and heat fluxes are calculated across ‘gates’ represented by the blue lines; the position of selected sites are labelled in bold: site radio soundings (RS), downstream part of the valley (DS), Mege`ve (MGV), Saint- Gervais-les-Bains (STGV), Chamonix (CHMNX) and Marnaz (MRNZ). (c) Terrain height along the polygon defining the PASSY control volume; the dotted line indicates the mean height of the terrain along the polygon; the dashed-dotted line is the height of the RS site; see text for details. (d) Snow height in the innermost domain d05 on 8 February 2015 at 1200 UTC; the position of selected sites are labelled in bold as in (b). conditions confirmed that the flow through the tributary can increase the along-valley mass flux by 5 to 10% compared to an idealised valley without a tributary. However, some of the complex interactions between the tributary flow and the down-valley flow were not captured with this idealised configuration. Coulter et al. (1989) showed that the mass flux through the tributary is extremely sensitive to the external flow. Under particular synoptic conditions the drainage from the tributary was found 20% larger than that expected if the mass flux was assumed proportional to the drainage area. Coulter et al. (1991) examined the mass fluxes through three tributaries of Kimball Creek, Colorado, and showed that the mass flux is largest for the tributary most closely aligned with the valley axis. Despite the efforts led in the ASCOT programme, large uncertainties remain about the effects of tributary flows on the heat and mass budget of a valley atmosphere. In this paper, we investigate the role of local and non-local controls, the latter including the synoptic wind above the valley boundary layer and flows from the surrounding tributary valleys affected by the large-scale flow, in the evolution of a PCAP using numerical model simulations. The PCAP event occurred in the section of the Arve River Valley around Passy in the French Alps, in February 2015, during the first intensive observation period (IOP1) of the PASSY-2015 field campaign (Staquet et al. 2015; Paci et al. 2016). To reach this aim, the mass and heat budgets in the Passy valley are analysed and compared to the ones for a semi-idealised simulation in which the synoptic flow in the initial condition is at rest. In Sect. 2, we present the design of the numerical simulations. Section 3 provides an overview of the synoptic conditions during IOP1. A detailed description of the episode and the evaluation of model data with field observations are reported in Sect. 4. Model results are analysed in Sect. 5 and 6. A summary and conclusions are presented in Sect. 7.This article is protected by copyright. All rights reserved. A cc ep te d A rti cl e 2. Methodology 2.1. The Passy Valley The Arve River Valley is located in the French Alps to the south of the Swiss Plateau, a relatively flat area located between the Jura mountain range and the Swiss Alps (see Fig. 1a). Hereafter the Passy Valley refers to the section of the Arve River Valley enclosed by the dashed black polygon in Fig. 1b. The mean height of the terrain along the polygon defining the PASSY control volume is zm = 1560 m a.s.l. (see Fig. 1c), and the highest massif surrounding the Arve River Valley is Mont-Blanc, culminating at 4808 m a.s.l. The valley-floor elevation decreases gradually down-valley from 588 m a.s.l. at Passy [upstream of the radio soundings site (RS)] to 488 m a.s.l. at Marnaz (MRNZ) over a distance of about 23 km. The width of the valley floor along this section of the valley increases gradually up-valley from just about 200 m upstream of MNRZ, where the valley presents a major constriction, to approximately 2000 m near the location of the RS site, and hardly decreases from there towards the escarpment leading to Chamonix (CHMNX). In the context of this work, the section of the Arve River Valley leading to CHMNX is considered a tributary of the Passy Valley. Therefore, the Passy valley presents three major tributaries leading to Mege`ve (MGV), Saint-Gervais- les-Bains (STGV) and CHMNX (see Fig. 1c). 2.2. Terrain representation and snow initialisation An accurate representation of the terrain is essential to numerical modelling in complex terrain such as alpine valley systems. Most alpine valleys are characterised by steep slopes that are a challenge for numerical models formulated using terrain- following coordinates, as is the case for the Weather Research and Forecasting (WRF) model used here. Even using a horizontal mesh size of the order of 3 km, slope angles exceed 25◦across the Alpine chain of western Europe. With a horizontal mesh size of 100 m, the maximum slope angle in the Arve River Valley is about 75◦. In the present work, 90-m terrain elevation data from the Shuttle Radar Topography Mission (SRTM, Farr et al. 2007) were interpolated at the horizontal mesh size of the nested model domains (see Fig. 1a). A smoothing filter was used to reduce the largest slope angles of the terrain to a maximum slope angle αmax = 45◦ while minimising any changes in the overall shape and small-scale features of the valley geometry. Several studies pointed out that the presence of snow on the ground surface alters significantly surface–atmosphere interactions (Za¨ngl 2005b; Tomasi et al. 2017; Neemann et al. 2015). The initialisation and modelling of the snow pack are difficult tasks. Numerical model analysis products are currently available at too coarse of a resolution (∼ 15 km) to be used to provide an appropriate representation of the snow cover at sub-kilometre scales. Furthermore, the initialisation of the snow density performed by the WRF Preprocessing System (WPS) assumes climatological values that are not representative for a specific case study. In the present work, MODIS/Terra (MOD10 L2) satellite products at a spatial resolution of 500 m (Hall et al. 2006) were used to initialise snow cover and albedo. The snow mask was directly interpolated at the horizontal grid points of the nested model domains. To initialise snow albedo, we followed the methodology outlined by Tomasi et al. (2017). Snow albedo is a function of the age of the snow (see Livneh et al. 2010), which is initialized as fresh snow (i.e. with the largest value for snow albedo, based on an analysis of climatological data, Robinson and Kukla 1985). A snow age corresponding to the satellite-retrieved albedo was calculated by inverting the functional relationship proposed by Livneh et al. (2010), which is implemented in the current version of the Noah land-surface Table 1. Spatial and temporal resolutions for the domains used for the real- case numerical simulations: number of points nx, ny and nz in the east, north and vertical directions, respectively, horizontal mesh size ∆x = ∆y, vertical grid spacing near the ground surface ∆zmin, time step ∆t and frequency fbu at which the lateral boundaries of the nested grids were updated. Domain nx, ny, nz ∆x = ∆y ∆zmin ∆t fbu d01 202, 202, 46 15 km 42 m 30 s 6 h d02 246, 246, 46 3 km 42 m 6 s 30 s d03 340, 340, 46 1 km 42 m 2 s 6 s d04 406, 406, 92 333 m 21 m 0.6 s 10 min d05 382, 382, 92 111 m 21 m 0.1 s 5 min model. The calculated snow age was then used as initial condition. The snow water equivalent was diagnosed by assuming that the density of the snow is, as a zero-order approximation, only function of the age of the snow (Meløysund et al. 2007). Hence, given the snow depth, the snow water equivalent can be specified at the initial time. Snow depth was assumed to depend only on the elevation of the underlying terrain, as in previous studies (e.g. Schmidli et al. 2009). Snow depth was initialised to increase linearly with height from a value of 0.05 m at sea level to a value of 0.15 m at 500 m above sea level (a.s.l.), and, from 500 m a.s.l., to a value of 1.5 m at 4500 m a.s.l. (see Fig. 1d). 2.3. Model setup The numerical simulations were performed using the WRF model with the advanced research core (ARW, Skamarock et al. 2008), version 3.5.1. Five nested domains were used (see Table 1 for the spatial and temporal resolutions that were used for the domains). The simulations were performed in three steps. Firstly, a simulation for the three outermost domains (d01, d02 and d03) was performed using online one-way nesting and a grid ratio of 5 and 3 for domains d02 and d03, respectively. The computations were made on 46 vertical layers. The grid was stretched along the vertical axis, with a vertical spacing of about 40 m near the ground surface (so that the first mass point above the surface is approximately at 20 m). The use of a relatively low vertical resolution in the outermost domains was required to keep both the model numerically stable and the simulated terrain height as close as possible to the real orography. Lateral boundary conditions for d02 and d03 were updated every parent domain time step. Initial conditions other than those for the snow fields and lateral boundary conditions for the outermost domain d01 were derived from the European Centre for Medium-range Weather Forecasts (ECMWF) gridded analyses available every 6 h with a horizontal mesh size of 0.25◦. The simulation was run continuously for 7 days from 7 to 14 February 2015. To minimise the deviation of the model results from the ECMWF analyses, spectral nudging was used for d01 and d02. Only the upper part of domain d03 (above 6000 m a.s.l.) was nudged towards the analyses. Snow cover and albedo were initialised from MODIS/Terra (MOD10 L2) satellite products (see Sect 2.2) averaged in time between 5 and 10 February 2015. Secondly, a simulation for domain d04 was performed using initial and lateral boundary conditions extracted from the results for d03, using offline nesting. The extent of d04 was chosen to cover the main regional-scale orographic feature surrounding the Passy Valley (see Fig. 1a). The horizontal mesh size was set to 333 m and the number of vertical levels increased to 92, compared to the outermost domains, with a vertical mesh size of about 20 m near the ground surface (so that the first mass point above the surface is approximately at 10 m) and 29 levels in the lowestThis article is protected by copyright. All rights reserved. A cc ep te d A rti cl e 1000 m above the surface. The lateral boundary conditions were updated every 10 min. This simulation was run continuously from 8 to 14 February 2015. Finally, a simulation for domain d05 was performed using a methodology similar to that used for domain d04. Differences are that the lateral boundary conditions were updated every 5 min in order to minimise spurious effects at these boundaries. Also, this simulation was reinitialised every day at 1200 UTC to facilitate the comparison with semi-idealised simulations (discussed below). The coverage of d05 is shown in Fig. 1b. The number of vertical levels was the same as that of d04 but the horizontal grid mesh size used was refined to 111 m. Results from additional sensitivity simulations revealed negligible differences when a domain size twice as large as that of d05 is used. Results from this sensitivity study also showed that a horizontal mesh size as fine as 111 m was required to resolve the constriction in the downstream part of the Passy valley. The constriction is closed for a coarser resolution which was found to significantly modify the dynamics of the near-surface valley atmosphere (not shown). The simulation for d05 is the reference simulation and is referred to as WRF-R hereafter. All nested domains were centred at the airport in Sallanches (45.935◦N, 6.636◦E) and shared the same ‘dynamics’ and ‘physics’ options, apart from that for the planetary boundary- layer (PBL) and subgrid-scale turbulence. A third-order Runge- Kutta scheme was used to integrate the model equations and a time-splitting technique was used to integrate the acoustic modes. The advection terms were discretised using a fifth-order Weighted Essentially Non-Oscillatory (WENO) scheme with positive definite filter. The PBL was parameterised for d01, d02 and d03 using the Yonsei University (YSU) parameterisation scheme (Hong 2010). In the higher-resolution simulations for d04 and d05 a 3D turbulent kinetic energy (TKE) diffusion scheme is used (and no PBL scheme), modified to enable the gravitational settling of fog, following the formulation proposed by Nakanishi (2000). Land surface processes were modelled using the Noah land surface model (Chen and Dudhia 2001) using four soil layers. A few modifications of the land surface model from the WRF model version 3.7 were implemented in the model used in the present work. These are the computation of the ground thermal conductivity, to prevent discontinous behavior during the melting of frozen water in the soil; and the calculation of the latent heat flux, to take into account dew deposition processes over snow covered surfaces. The Monin-Obukhov similarity theory (Jime´nez et al. 2012) is used to couple the land-surface to the atmosphere. Shortwave and longwave radiation were parameterised using the Rapid Radiative Transfer Model (Mlawer et al. 1997). Shadowing effects were included to take into account the effect of the mountains on the radiative fluxes. Microphysics was parameterised using the scheme developed by Morrison et al. (2005), modified for the treatment of ice fog following Neemann et al. (2015). The number of cloud condensation nuclei was increased to 500× 106, a value more representative of continental environments. A set of semi-idealised numerical simulations was performed to extract the effects of the large-scale flow (simulation WRF-I). WRF-I differs from WRF-R in that the domain is twice as large (see Fig. 1a), and that at the initial time the temperature and humidity fields are horizontally homogeneous and the wind speed is set to zero. The use of a larger domain in WRF-I compared to WRF-R was necessary to minimise spurious effects at the lateral boundaries, at which the normal velocity is set to zero. Hence, the main difference between WRF-I and WRF-R lies in the absence of large-scale flow for WRF-I. The latter simulation was run for the period between 10 February 2015 at 1200 UTC and 12 February 2015 at 1200 UTC and re-initialised on 11 February 2015 at 1200 UTC. The initial temperature and humidity profiles for WRF-I were extracted from the real-case simulation for d04 at the centre of the domain at the time at which WRF-I was initialised, namely 10 and 11 February 2015 at 1200 UTC. The skin temperature was initialised by extrapolating the temperature of the first three air layers above the ground surface. 3. Overview of the large-scale circulation during IOP1 The PCAP that formed in the Passy Valley during IOP1 was the result of the passage of an upper level ridge over the region. The geopotential height at 500 hPa showed a dipole pattern over Europe between 2 and 8 February 2015, characterized by an upper level ridge adjacent to a trough. The upper level ridge of the dipole pattern was located immediately to the west of the British Isles on 9 February 2015 at 0000 UTC (see Fig. 2a). The PCAP formed in the evening on 9 February 2015, when the upper level ridge moved over Northwest Europe (see Fig. 2b). This synoptic flow pattern led to the advection of warmer air above the western Alps, setting the conditions for the PCAP to develop in the Passy Valley. From 10 to 12 February 2015 the upper level ridge continued to move across Europe and was centred over Central Europe on 12 February 2015. The displacement of the ridge resulted in a weakening of the wind speed and a veering of the wind with time over the western Alps from north-easterly on 9 February 2015 to south-easterly on 12 February 2015 (see Fig. 2b to d), allowing the CAP to persist over time. On 13 February 2015 the trailing upper level trough moved across Northwest Europe, leading to a southerly flow and an increase in wind speed over the western Alps (see Fig. 2e). On 14 February 2015 a large-scale perturbation reached the western Alps (see Fig. 2f), thereby removing the PCAP and putting an end to IOP1. 4. Life-cycle of the persistent cold-air pool during IOP1 4.1. Vertical structure of the cold-air pool Figure 3 shows the temporal evolution of the vertical structure of the PCAP obtained by compiling potential temperature profiles from the radiosonde ascents and wind profiles retrieved from LiDAR and wind profiler data at the RS site during IOP1, and extracted from the WRF-R simulation. The temporal evolution is overall well captured in WRF-R. However discrepancies can be noted, in particular before the PCAP event during the night from 8 to 9 February 2015 and at the end of the episode on 14 February 2015 when the simulated temperature and wind speed are overestimated above zm. Vertical profiles of the root-mean- squared-error (RMSE) and bias of potential temperature and wind speed between 9 February 2015 at 1200UTC to 14 February 2015 at 1200UTC are also displayed to quantitatively evaluate the numerical results of the simulations. Results before 9 February at 1200UTC were not included in the computation of the statistics, because the simulation in the outer domain d04 was not fully spun up yet for these times. Statistics for the wind direction are not displayed as at low wind speeds, the fluctuations of this quantity are large, leading to large RMSE values (Chow et al. 2006). The RMSE of the potential temperature averaged between the valley floor and zm is about 1.6 K, decreasing above zm to about 1 K. The value of the bias is−0.32 K averaged between the valley floor and zm, indicating a slight systematic underestimation of the potential temperature in the simulated PCAP above the valley floor. This statistical analysis indicate that the model is able to represent the temporal variability of the observed temperature. Regarding the wind speed, the model shows a nearly zero bias compared to observations between the valley floor and zm, whereas the RMSE is in between 1 and 2 m s−1.This article is protected by copyright. All rights reserved. A cc ep te d A rti cl e Figure 2. Synoptic weather conditions from 9 to 14 February 2015 at 0000 UTC (a) to (f) extracted from the European Centre for Medium-range Weather Forecasts (ECMWF) gridded analyses with a horizontal grid point spacing of 0.25◦. The colour contours indicate geopotential height at 500 hPa and wind barbs indicate the wind at 700 hPa. The green dot in the centre of the weather charts indicates the position of the Passy Valley. The PCAP formed in the evening on 9 February 2015, when warm air was advected above the Passy Valley as the upper level ridge approached the western Alps (see Sect. 3). Note the weakening and veering of the flow associated with the passage of the ridge (see Fig. 3c to f and Fig. 2). The formation of the PCAP led to a rapid increase of the vertically-integrated valley heat deficit (VHD, Whiteman et al. 1999) from 3 to 12 MJ m−2, which is well reproduced in WRF-R (see Fig. 4a). Between 9 February 2015 at 1200 UTC and 12 February 2015 at 1200 UTC, the structure of the PCAP in the Passy Valley varied substantially, while the wind speed below 2500 m a.s.l. was in the order of a few metres per second. During this period the height of the PCAP decreased gradually with time, and so the VHD. This decrease was most rapid during the night from 10 to 11 February 2015, by 4–5 MJ m−2. Despite the VHD kept decreasing with time, the nighttime near-surface potential temperature gradient computed over the lowest 100 m above the valley floor, ∆θ/∆z100, was found to increase from day to day (see Fig. 4b). This behaviour is well reproduced in WRF-R, but with an underestimation of its amplitude. The different behaviour of VHD and ∆θ/∆z100 suggests that a thermal decoupling between the near-surface atmosphere and the atmosphere aloft occurs, which will be analysed in details in the following Sections. The maximum value of ∆θ/∆z100 was found on 12 February 2015 at 0000 UTC, as a result of the continuous advection of warm air within the valley atmosphere below z = zm. On 13 February 2015 the wind speed in the valley atmosphere increased above 5 m s−1 as a large-scale perturbation approached the western Alps (see Sect. 3). On 14 February 2015 the wind speed increased also at the valley bottom, leading to mixing and the full destruction of the PCAP. 4.2. Regional-scale circulation A more detailed account of the interplay between the local (valley- scale) flow and the regional-scale flow at z = zm, simulated in domain d04, is given in Fig. 5. Timeseries of wind and temperature at selected sites in the Passy Valley, RS and DS, and in the main tributary valleys, MGV and CHMNX (see Fig. 1b for the location of the sites), are displayed in Fig. 6 to show the response of these fields (in WRF-R) to the large-scale flow. On 10 February 2015 at 0000 UTC, the north-easterly large- scale flow was channelled through the downstream part of theThis article is protected by copyright. All rights reserved. A cc ep te d A rti cl e Figure 3. Temporal evolution of the vertical structure of the cold-air pool as obtained from the WRF-R simulation (left column) and from observational data (right column). (a) and (b): potential temperature; root-mean-squared-error and bias of the potential temperature (c) and (d): wind speed (ws), (e) and (f): wind direction. The potential temperature data are obtained from the radiosonde (RS) ascents, the wind speed and wind direction are retrieved from the LiDAR (below 1050 m a.s.l.) and the wind profiler (above 1050 m a.s.l.) at the RS site during IOP1. The dashed lines indicates zm = 1560 m a.s.l. (see Fig. 1c). The RS site is shown in Fig. 1b. The middle panels display the root-mean-squared-errors (continuous line) and bias (dashed line) of potential temperature (top row) and wind speed (middle row), with respect to observations. Figure 4. Timeseries of the valley heat deficit, vertically integrated from the ground surface to the mean height of the terrain zm = 1560 m a.s.l. (a), and of the near-surface potential temperature gradient over the lowest 100 m above the surface (b), at the RS site (see Fig. 1b), derived from data from the radiosonde ascents and computed from the WRF-R and WRF-I outputs. This article is protected by copyright. All rights reserved. A cc ep te d A rti cl e Fi gu re 5. H or iz on ta lm ap of th e w in d (v ec to rs ), po te nt ia lt em pe ra tu re (c ol ou re d co nt ou rs )a nd te rr ai n he ig ht (b la ck lin e co nt ou rs )fi el ds in do m ai n d0 4 at z = z m = 15 60 m a. s. l. on 10 Fe br ua ry 20 15 at 00 00 U T C an d 12 00 U T C (a )a nd (b ), re sp ec tiv el y, 11 Fe br ua ry 20 15 at 00 00 U T C an d 15 00 U T C (c )a nd (d ), re sp ec tiv el y, 12 Fe br ua ry 20 15 at 12 00 U T C (e )a nd 14 Fe br ua ry 20 15 at 00 00 U T C (f ). A re as fo rw hi ch th e te rr ai n he ig ht is ab ov e th e di sp la ye d he ig ht le ve lz = z m ar e m as ke d ou tt o be tte rv is ua lis e th e flo w ch an ne lin g. T he sp ee d an d di re ct io n of th e la rg e- sc al e flo w at z = 60 00 m a. s. l. av er ag ed ov er th e lo ca tio ns of th e R S, D S, M G V an d C H M N X si te s (s ee Fi g. 1b fo r th e lo ca tio n of th e si te s) ar e in di ca te d on th e to p ri gh tc or ne r of th e pl ot s. T he gr ee n do ts in th e ce nt re of ea ch pa ne li nd ic at e th e po si tio n of th e Pa ss y V al le y. A zo om on do m ai n d0 5 is pr ov id ed in th e bo tto m ri gh tc or ne ro ft he pl ot . This article is protected by copyright. All rights reserved. A cc ep te d A rti cl e Passy Valley, and split between CHMNX and MGV at z= zm (see Fig. 5a and 6e). Warm air advection occurred over night across the entire region, where potential temperature θ at z = zm increased by 7 K in 12 hours (see Fig. 6f), while it decreased by 2–3 K at z= 1000 m a.s.l. (see Fig. 6i). This process, that is warming at mid levels and cooling at lower levels, led to the initial strengthening of the PCAP within the Passy Valley, and it is well represented in the numerical simulation. At 1200 UTC, the large-scale flow rotated to a south-westerly direction, reaching a minimum wind speed of less than 1 m s−1 at the different sites (see Fig. 6a and 6b). A strong and deep down-valley (north-easterly) wind formed in the Rhone Valley, located north-east of the Chamonix Valley near Sion, which was channelled through the gap that connects the two valleys, strengthening the down-valley (north-easterly) flow in CHMNX (see Fig. 5b). Potential temperature θ at z = zm continuously increased on 10 February 2015, reaching a maximum value at 2100 UTC (see Fig. 6f), that is when the upper level ridge just passed over the Passy Valley. This can be inferred by noticing the rotation and weakening of the wind aloft (at z = 6000 m a.s.l., see Fig. 6b and Fig. 6a, respectively). In the early night between 10 to 11 February 2015 the near-surface flow from MGV reversed to down-valley (south-westerly) as the upper level ridge moved away from the western Alps, and remained down-valley (south-westerly) until the end of the simulated time period (see Fig. 6k). During the night between 10 to 11 February 2015 θ at z = 1000 m a.s.l. increased, while it stayed approximately constant at z = zm (see Fig. 6f and i). The large-scale flow at this time was northeast-easterly, i.e. aligned with the Rhone and Chamonix valleys, thereby favouring the channelling of the flow down these two valleys (see Fig. 5c). Throughout the night the large-scale flow slowly rotated clockwise and strengthened as the upper level ridge moved towards the north-east (see Fig. 6a and 6b). Later on, on 11 February 2015 at 1200 UTC, the large- scale flow veered to south-easterly, and the resulting channelling through the Swiss Plateau gave rise to a regional south-westerly flow at mid-level (see Fig, 5d), leading to a channelling of the flow through the downstream part of the Passy Valley. During the night between 11 and 12 February 2015 θ at z = 1000 m a.s.l. and z = zm hardly varied (Fig. 6f and i), while the near- surface temperature dropped throughout the night, as during the previous night (Fig. 6l). During this night the amplitude of the diurnal range of θ at z = zm is underestimated in WRF-R, but the overall temporal evolution is well captured. From 10 to 12 February 2015 the near-surface temperature at night was non-homogeneous across the valley floor, as indicated by differences of 4–5 K between the DS and RS sites that are approximately at the same height (see Fig. 6l), while the wind speed there was very weak (less than 1 m s−1 on average during this time period). It is worth noting that simulated near- surface temperatures are slightly underestimated at night, though they fall within the observed range when spatially averaged within a radius of a few hundred metres around the sites (not shown). A cold bias of about 3 K is also present in the afternoon. The errors in the representation of the amplitude of the diurnal cycle of the near-surface temperature over snow-covered surfaces in anticyclonic conditions can be related to issues in the representation of turbulent surface fluxes and conductive heat flux into the snowpack. Between 10 February at 1200UTC and 12 February at 1200UTC, the RMSE of near-surface potential temperature, wind speed and direction are 2.9 K, 0.3 m s−1 and 65◦, respectively. The values for the potential temperature are comparable or smaller than the ones reported in previous modeling studies of CAPs and PCAPs (see for instance Page`s et al. 2017). It should be noted that the RMSE of the wind speed and direction are computed on a smaller number of data. As before, the relatively large value of the RMSE of the wind direction can be related to the large fluctuations of this quantity occurring at low wind speeds. During 13 and 14 February 2015, advection of cold air by the large-scale perturbation led to the destruction of the PCAP, strengthening of the near-surface flow with prevailing southerly winds, and a homogenisation of the temperature across the valley floor (see Fig. 5g, 5h and 6l). The overview of the life-cycle of the PCAP presented above leads to the identification of three main time periods, which correspond to the three stages of a PCAP: P1 The formation stage – from 8 to 10 February 2015, when the temperature inversion formed in the Passy Valley and a change in wind direction was observed aloft; P2 The persistent stage – from 10 to 13 February 2015, when the temperature inversion persisted in the Passy Valley; P3 The destruction stage – on 14 February 2015. Since the large-scale flow shows substantial variations with time in wind speed and direction throughout the anticyclonic period, the persistent stage is divided in two sub-periods: P2a From 10 February 2015 at 1200 UTC to 11 February 2015 at 1200 UTC, when the upper level ridge moved over the Passy Valley, and the wind speed at z = 6000 m a.s.l. increased over night from a minimum value; P2b From 11 February 2015 at 1200 UTC to 12 February 2015 at 1200 UTC, when the large-scale flow was veering from south-easterly to southerly, and the wind speed at z = 6000 m a.s.l. hardly varies with time. Because the vertical structure and bulk properties of the PCAP show a different behaviour during the two sub-periods of the persistent stage, the link with the variations of the large-scale flow during these periods is analysed in the following sections. 4.3. Valley-scale circulation during the persistent stage In this section the valley-scale circulation during the persistent stage P2 is described by comparing results from WRF-R and the semi-idealised simulation WRF-I. A mass and heat budget analysis and a mechanistic description of the flow behaviour presented in this section will be discussed in Sect. 5. 4.3.1. Sub-period P2a Sub-period P2a (from 10 February 2015 at 1200 UTC to 11 February 2015 at 1200 UTC) is characterised by a maximum VHD around 1800 UTC, followed by a gradual decrease of the VHD during the course of the night. The wind speed near the surface and at z = 1000 m a.s.l. was at most 1 m s−1 at the different sites considered in Fig. 6, except MGV, where it was maximum and in the order of 3–4 m s−1 near the surface during P2a. The non-local control of the large-scale flow on the near- surface circulation can be extracted by comparing results from WRF-R and WRF-I (see Fig. 7). Between 1800 UTC and 0600 UTC on the following day, the circulation at the bottom of the Passy Valley is qualitatively similar in WRF-R and WRF-I. There is a clear down-valley flow in the downstream part of the valley and down-valley flows from CHMNX and STGV mostly detach above the valley floor. At the valley exit the along-valley flow is found to accelerate because of the narrowing of the valley. The along-valley flow in MGV is up-valley at 1800 UTC both in WRF-I and WRF-R, and reverses to down-valley at the surface during the night (see Fig. 7b and d). However this flow is weaker in WRF-I than in WRF-R, with wind speeds of about 0.3–0.5 m s−1 and 3–4 m s−1, respectively.This article is protected by copyright. All rights reserved. A cc ep te d A rti cl e P2 a P2 b P2 a P2 b P2 a P2 b Fi gu re 6. Ti m es er ie s of th e ho ri zo nt al w in d sp ee d (l ef tc ol um n) an d di re ct io n (m id dl e co lu m n) ,a nd po te nt ia lt em pe ra tu re (r ig ht co lu m n) at z = 60 00 m a. s. l., z = z m = 15 60 m a. s. l., z = 10 00 m a. s. l. an d z = 10 m a. g. l. (t op to bo tto m ro w s, re sp ec tiv el y) , at th e lo ca tio ns of th e R S, D S, M G V an d C H M N X si te s (s ee Fi g. 1b fo r th e lo ca tio n of th e si te s) ,i n W R F- R du ri ng th e si m ul at ed tim e pe ri od .O bs er va tio na ld at a at th e R S si te ar e su pe ri m po se d (b la ck do ts ): w in d da ta de riv ed fr om th e w in d pr ofi le r m ea su re m en ts at z = z m an d z = 10 00 m a. s. l. an d fr om an in st ru m en te d m et eo ro lo gi ca l m as t at z = 10 m a. g. l., an d po te nt ia l te m pe ra tu re da ta at z = z m an d z = 10 00 m a. s. l. de riv ed fr om th e ra di os on de m ea su re m en ts an d fr om an in st ru m en te d m et eo ro lo gi ca lm as ta tz = 10 m a. g. l.. Si m ul at ed da ta at th e R S si te ar e di sp la ye d w ith a th ic k bl ac k lin e to fa ci lit at e co m pa ri so n w ith ob se rv at io ns . This article is protected by copyright. All rights reserved. A cc ep te d A rti cl e Figure 7. Simulated wind and potential temperature fields at z= 10 m a.g.l. in WRF-R (top row) and WRF-I (bottom row) on 10 February 2015 at 1800 UTC (left column) and 11 February 2015 at 0600 UTC (right column). The speed and direction of the large-scale flow at z = 6000 m a.s.l. are indicated on the top right corner of the plots for WRF-R. The discontinuity in the magnitude of the wind vectors in Fig. 7b where the air flows from MGV to the Passy Valley indicates that the flow detaches from the surface because of the strong stratification of the PCAP. This detachment is clearly visible in the composite vertical cross-sections that approximately follows the flow (along the white line in Fig. 1b), shown in Fig. 8. The capping inversion at the top of the PCAP lowers down progressively and ‘sharpens’ (i.e. its depth decreases) as the flow through MGV strengthens during the course of the night (see Fig. 8a,c and e). A hydraulic jump forms as the air descends into the Passy Valley (see Fig. 8e). Conversely, although the flow from MGV in WRF-I reverses above the near-surface layer only around 0000 UTC, the upper part of the PCAP cools steadily throughout the night (see Fig. 8b,d and f), resulting in a deeper PCAP by the end of the night compared to WRF-R. The differences in the flow from the tributaries between WRF-R and WRF-R explain the different behaviour of the VHD in the two simulations during the night of P2a (see Fig. 4a). In WRF-R, the lowering of the PCAP height due to the flow through the tributaries leads to the reduction of VHD throughout the night, whereas in WRF-I the VHD remains about constant with time due to the steady cooling of the atmospheric column. The lower part of the PCAP displays a temporal evolution of wind speed and thermal structure similar to that in WRF-R, as can be inferred from the isentrope θ = 276 K in Fig. 8. This result suggests that the flow from MGV mostly controlled the increase in temperature around z = 1000 m a.s.l. between 0000 and 0600 UTC, and more generally the thermal structure of the upper part of the PCAP, while the lower part of the PCAP was mostly controlled by local processes. This will be quantified in Sect. 5, where the heat and mass fluxes through the tributaries of the Passy Valley atmosphere are analyzed. At 0600 UTC, the near-surface flow in the part of the Passy Valley upstream of RS is weak and almost stagnant in both WRF-R and WRF-I (see Fig. 7b and 7d). However, the near-surface temperature there is lower in WRF-I than in WRF-R, which explains the larger near-surface vertical gradient of potential temperature that develops in WRF-I when compared to WRF-R during the second part of the night (see Fig. 4b). 4.3.2. Sub-period P2b During the sub-period P2b (from 11 February 2015 at 1200 UTC to 12 February 2015 at 1200 UTC) the evolution of the bulk thermal properties of the Passy Valley atmosphere are more similar in WRF-R and WRF-I than during P2a (see Fig. 4). At 1800 UTC, the near-surface flows through MGV, CHMNX and STGV are down these tributaries and the near-surface flow in the Passy Valley is down-valley (see Fig. 9a and 9c). Hence, the near- surface flow and the flow in the upper part of the PCAP are in opposite directions, i.e. they are dynamically decoupled (see Fig. 5 and Sect. 4.2). The speed of the flow from MGV increases with time in WRF-R, reaching a maximum value in the morning on 12 February 2015, while it remains constant in WRF-I (see Fig. 9b and 9d). The large scale flow slightly rotates during the night, being more aligned with the MGV valley axis in the morning, suggesting that the increase of wind speed at MGV is an effect of the channelling of this large scale flow. The flow from CHMNX is weaker than that from MGV and comparable in WRF-R and WRF-I. In contrast to P2a, the near-surface temperature in theThis article is protected by copyright. All rights reserved. A cc ep te d A rti cl e Figure 8. Composite vertical cross-sections following the flow (along the white line in Fig. 1b) of horizontal wind speed (in the plane of the cross-section) and potential temperature (K) in WRF-R (left column) and WRF-I (right column) on 10 February 2015 at 1800 UTC, and 11 February 2015 at 0000 and 0400 UTC (top to bottom rows, respectively). Passy Valley is similar in WRF-R and WRF-I during P2b (see also Fig. 4b). 5. Mass and heat fluxes through the Passy Valley Sect. 4.3 highlighted non-local controls of the large-scale flow on the PCAP in the Passy Valley during the persistent stage (P2), in particular in determining the respective contributions of the tributary flows to the mass and heat budgets of the valley atmosphere. In this section, these contributions are quantified. 5.1. Mass fluxes The mass budget of the Passy Valley atmosphere is performed over the PASSY control volume, defined in Sect. 2.1. Assuming incompressibility, the mass budget reads ∑ i ∫ Ai ρ vi ni dS︸ ︷︷ ︸ Mi = MDS +MMGV +MCHMNX +MSTGV +MTOP = 0, (1) This article is protected by copyright. All rights reserved. A cc ep te d A rti cl e Figure 9. Simulated wind and potential temperature fields at z= 10 m a.g.l. in WRF-R (top row) and WRF-I (bottom row) on 11 February 2015 at 1800 UTC (left column) and 12 February 2015 at 0600 UTC (right column). The speed and direction of the large-scale flow at z = 6000 m a.s.l. are indicated on the top right corner of the plots for WRF-R. (a) (b) Figure 10. Timeseries of the mass fluxes through DS (MDS), MGV (MMGV), CHMNX (MCHMNX), STGV (MSTGV), vertically integrated from the ground surface to z = zm, and through the top surface at z = zm (MTOP) for the period from 10 February 2015 at 1200 UTC to 12 February 2015 at 1200 UTC in WRF-R (a) and WRF-I (b); see Fig. 1b and 1c for the location of the gates. The nighttime periods (from sunset to sunrise) are displayed as grey shading. The bottom panel shows the mass fluxes normalised by the sum of all mass fluxes out of the Passy Valley (i.e. a positive quantity). Positive and negative values of the mass flux correspond to a flux out of and into the Passy Valley, respectively. Note the different scale used between panel (a) and (b). where Ai = ADS, AMGV, ACHMNX, ASTGV, are the lateral surfaces of the control volume (see the blue surfaces in Fig. 1c), Ai = ATOP is the top surface at z = zm, ρ is the air density, vi is the wind component normal to the surface Ai orientated by the unit vector ni (defined positive outwards). Mi = MDS, MMGV, MCHMNX, MSTGV are the mass fluxes across the surface Ai. To compute the massThis article is protected by copyright. All rights reserved. A cc ep te d A rti cl e fluxes, the model fields have to be interpolated on a cartesian grid. For the vertical mass flux, this can lead to a misrepresented contribution from the slope flows from the terrain surrounding the PASSY control volume. Hence, to take into account these effects, in the following the vertical mass flux is computed as the residual of the mass budget. 5.1.1. Sub-period P2a As one may expect, DS is the main exit for the air mass within the valley (see Fig. 10a). During P2a, both CHMNX and STGV contribute an inflow into the valley, while the mass flux through MGV can be in or out of the valley. When averaged during nighttime (from sunset to sunrise), MCHMNX and MSTGV account for 35% and 18% of the total flow of mass out of the valley, respectively. When combined together, the flow from these tributaries provide about half of the mass flowing out of the valley at night during P2a. The other half is coming from MMGV and/or MTOP. MMGV > 0 (outflow) when the upper level ridge passed over the valley (i.e. when the wind speed aloft decreased to its minimum value, see Fig. 6a), and subsiding motions at the valley top account for about 40% of the inflow of mass into the valley. As the upper level ridge moves away from the region (and the wind speed increases at upper level, see Fig. 6a), most of the air mass out of the valley flows through DS, and MTOP decreases continuously while the flow through MGV reverses (MMGV changes from positive to negative at about 2200 UTC). In the late night from 0300 to 0600 UTC, the inflow of mass from MGV exceeds that from CHMNX and STGV and the total mass originating from the tributaries exceeds that flowing through DS, resulting in an export of mass across the top of the valley volume. The mass fluxes in WRF-I are one order of magnitude smaller than those in WRF-R (see Fig. 10b). The nighttime-averaged contribution to the total flow of mass out of the valley from CHMNX is smaller in WRF-I than in WRF-R (25% compared to 35%), implying that the down-valley flow at CHMNX in WRF- R is affected by the channelling of the flow from the Sion valley. Overall, the mass flux analysis demonstrates that the large-scale flow controls the magnitude of the tributary flows and allows to quantify the relative contribution of the latter flows to the total outflow from the Valley during P2a. 5.1.2. Sub-period P2b During P2b, between 1300 and 2000 UTC the large-scale flow was channelled over the Swiss Plateau (see Fig. 6b and Fig. 5d), resulting in a strong inflow into the downstream part of the Passy Valley (see Fig. 10a). This inflow through DS is mostly balanced by the outflow through MGV and the export of mass across the top of the valley volume. MDS reverses from negative to positive at 2000 UTC when the large-scale flow rotates to south-southeasterly and so was in the down-valley direction. At this time, MMGV > 0 and MTOP, MCHMNX and MSTGV contribute equally to the total drainage out of the valley through DS and MGV. As the large- scale flow rotated to southerly, the flow at MGV reverses from up- to down-valley and strengthens, leading to a continuous increase of MMGV after 0000 UTC. MCHMNX decreases during the course of the night, contributing on average 15% of the total drainage out of the valley, compared to 35% during P2a. Conversely, the nighttime-averaged contribution to the total flow of mass out of the valley from STGV is 16% and comparable to that during P2a. This suggests that the flow through STGV is less sensitive to variations in the large-scale flow than those in CHMNX and MGV. In WRF-I, each tributary contributes about 30% of the total drainage out of the Passy Valley through DS during P2b. Interestingly MMGV < 0 throughout P2b, as opposed to P2a, implying that the vertical structure of the PCAP influences the direction of the flow through MGV. This can be explained as follows: since the PCAP at 1200 UTC is much shallower than during P2a, and its top is below the height of MGV, there is no horizontal temperature contrast between RS and MGV and hence no thermally-driven up-valley flow from the Passy Valley to MGV. 5.1.3. Vertical and horizontal structure of mass fluxes The analysis presented so far is further developed by examining the vertical structure of the mass fluxes through DS. A significant fraction of the total drainage out of the Passy Valley occurs in the upper part of the PCAP as a result of the detachment of the flow from the tributaries (see Fig. 11a). The mass flux through DS in the lower part of the PCAP is more variable and has a similar magnitude in WRF-R and WRF-I. This suggests that despite the flow from the tributaries is a major control on the total outflow from the valley, it mainly affects the upper part of the PCAP with only an indirect effect on the lower part. Indeed, the mass flux in WRF-R is characterized by temporal oscillations between outflow (positive) and inflow (negative) in WRF-R. Such oscillations are also visible in WRF-I but only during P2b (see Fig. 11b). These oscillations can be triggered either by those of thermally-driven flows or by internal gravity waves (IGWs) created by the detachment of the flow through the tributaries at the top of the CAP via a hydraulic jump (see for instance Largeron et al. 2013). A closer inspection of Fig. 11a and 11b reveals that the slope of the contours associated with the oscillations with respect to the vertical is non zero, indicating that there is a phase propagation in the vertical. As a consequence, the hypothesis that oscillations could be triggered by those of thermally-driven flows can be rejected, as those are pure oscillations in time with no phase propagation (Largeron et al. 2013; Quimbayo-Duarte et al. 2019). The horizontal cross-section of the vertical mass flux at the height of STGV during P2a at 2200UTC confirms the IGW mechanism behind the mass flux oscillations in WRF-R (see Fig. 12a). In WRF-R the flow through STGV detaches at about z= 1000 m a.s.l., leading to the formation of internal gravity waves that propagate within the Passy Valley atmosphere. In WRF-I, internal gravity waves are excited by the detachment of the flow from MGV and propagate downstream (the link with the waves observed at DS needs to be clarified) (see Fig. 12b). A different time period is selected for WRF-I compared to WRF-R because the tributary flows are characterised by a different magnitude and time of occurence in the two simulations. 5.2. Heat fluxes and impact on the CAP height The previous subsection revealed the key role of the inflows/outflows from the tributary valleys in the mass budget of the Passy Valley atmosphere during the persistent stage of the CAP and how those flows are affected by the large-scale flow. This section examines their role on the heat budget of the Passy Valley atmosphere. The heat fluxes are computed from the non-linear advective term of the volume-averaged density-weighted heat budget equation over the PASSY control volume, that is, assuming incompressibility, ADV = 1 [Mass]P ∫ V ρ v j ∂θ ∂x j dV =∑ i 1 [Mass]P ∫ Ai ρ ( θ ′ v j ) n j dS = HDS +HMGV +HCHMNX +HSTGV +HTOP, (2) where [Mass]P is the mass of the PASSY control volume, θ ′ = θ − θ and θ is the volume-averaged potential temperature inThis article is protected by copyright. All rights reserved. A cc ep te d A rti cl e Figure 11. Time-height plot of the mass flux through the downstream (DS) gate for the period from 10 February 2015 at 1200 UTC to 12 February 2015 at 1200 UTC in WRF-R (a) and WRF-I (b); see Fig. 1b and 1c for the location of the gate. The nighttime periods (from sunset to sunrise) are displayed as dashed vertical lines. Note the different scale used between panel (a) and (b). Figure 12. Horizontal cross-sections (across the polygon defining the PASSY volume, see Fig. 1b) of the vertical mass flux at the height of STGV on 10 February 2015 at 2200 UTC (during P2a) in WRF-R (a) and at the height of MGV on 11 February 2015 at 1900 UTC (during P2b) in WRF-I (b). Positive and negative values of the mass flux correspond to a flux out of and into the Passy Valley, respectively. Note the different scale used between panel (a) and (b). the PASSY volume (see for instance Lee et al. 2004), and Ai the lateral and top surfaces of the volume, defined in Sect. 5.1. Hi = HDS, HMGV, HCHMNX, HSTGV are the heat fluxes across the surface Ai, divided by [Mass]P. The heat flux through the top surface is computed as a residual of Eq. 2, as it is done for the vertical mass flux (see Sect. 5.1). In order to quantify the contribution of advective processes to the heat budget of the valley atmosphere, the ADV term above is compared with the volume- averaged density-weighted total tendency, that is TEND = 1 [Mass]P ∫ V ρ ∂θ ∂ t dV . (3) TEND is the variation of the heat content of the PASSY volume due to the advection of heat through the lateral and top surfaces of the volume, as expressed in Eq. 2, and the vertical divergence of radiative and turbulent heat fluxes. As clear-sky conditions persisted through P2 in WRF-R (expect for some sporadic fog that formed near the surface during the night of P2a, not shown), latent heat release/absorption associated to phase changes can be neglected in the heat budget at the valley scale. The effect of the heat fluxes through the tributaries on the thermal structure of the PCAP is evaluated in WRF-R by investigating the temporal evolution of the CAP height, CAPh, and the bottom height of the capping inversion, CIh. CAPh is defined as the maximum height where the vertical gradient of absolute temperature is positive and CIh is defined as the height where the curvature of θ (z) presents a positive maximum. During P2, when a capping inversion layer exists within the valley atmosphere, CAPh coincides with the top height of this layer, so that the depth of the capping inversion d can be defined as d = CAPh−CIh. 5.2.1. Sub-period P2a In WRF-R the temporal evolution of TEND during P2 follows that of the advection contribution ADV (see Fig. 13a). TEND is controlled by ADV during P2a while it is close to zero during P2b (implying a balance between ADV and the contributions from turbulent heat and radiative flux divergences). The large variations in ADV during P2a are associated with the variability of the advection contributions from the tributaries and top surface of the valley volume. Conversely, in WRF-I the total tendency and advection term have opposite signs and a steady temporal evolution (see Fig. 13c). Advection contributes to a warming ofThis article is protected by copyright. All rights reserved. A cc ep te d A rti cl e (c) (d) (a) (b) Figure 13. (a) Timeseries of the volume-averaged density-weighted total potential temperature tendency (TEND), advection contribution (ADV) to the total tendency and that restricted to the tributaries (Trib, horizontal advection), averaged over the PASSY volume (see Fig. 1b), for the period from 10 February 2015 at 1200 UTC to 12 February 2015 at 1200 UTC in WRF-R; (b) Timeseries of the horizontal advection contribution to the total tendency from DS, MGV, CHMNX and STGV in WRF-R (see Fig. 1b). (c) and (d): same as (a) and (b), but for WRF-I. Note the different scale used for the y-axis between panels (a), (b) and panels (c), (d). The nighttime periods (from sunset to sunrise) are displayed as grey shading. (b)(a) Figure 14. (a) Timeseries of the heights of the cold-air pool (CAPh) and of the capping inversion (CIh), and potential temperature at the RS and MGV sites (see Fig. 1b), vertically-averaged in the height range 1150 < z < zm = 1560 m a.s.l (θRS and θMGV, respectively), for the period from 10 February 2015 at 1200 UTC to 12 February 2015 at 1200 UTC in WRF-R. The nighttime periods (from sunset to sunrise) are displayed as grey shading. Note that CIh subsides down to the valley floor after 11 February 2015 at 1500UTC; see the text for more details. (b) Vertical profiles of potential temperature at the RS site for selected times during P2a and P2b in WRF-R: 10 February 2015 at 1230 UTC (continuous red line), 10 February 2015 at 1930 UTC (dashed red line), 11 February 2015 at 0430 UTC (dotted red line) and 11 February 2015 at 0900 UTC (dashed-dotted red line). In both panels, dashed lines indicate the height of the valley floor at MGV (1100 m a.g.l.) and zm = 1560 m a.s.l. (see Fig. 1c). the valley atmosphere during most of the analysed time period, a behaviour that is typical in synoptically unperturbed conditions and reported in previous studies of idealised CAPs (Schmidli and Rotunno 2010; Arduini et al. 2017), hence demonstrating the impact of the large-scale flow on the valley heat budget in WRF-R. A more detailed analysis of the advection contribution in WRF-R reveals the large contribution of tributary flows. In general, the sum of the heat fluxes through MGV, STGV and CHMNX is mostly balanced by an opposite heat flux through DS (see Fig. 13b). However, the (un)balance between these fluxes varies with time and responds to variations in the large-scale flow, resulting in a net heat input (warming) or export (cooling) from the tributaries, with a large impact on the evolution of the thermal structure of the PCAP. In WRF-I, there is a close balance between the horizontal advection through DS and the vertical advection, with only a minor contribution from the other tributary valleys (see Fig. 13b). During the early night until 2200 UTC, the flow through MGV is out of the valley volume (see Fig. 10a) and PCAPh decreases from z = zm to z = 1300 m a.s.l (see Fig. 14a). This is mainlyThis article is protected by copyright. All rights reserved. A cc ep te d A rti cl e attributed to the warming of the upper part of the PCAP due to subsiding motions across the top surface of the valley volume (70%) and ADV through the other two tributaries, CHMNX and STGV (30%); see Fig. 13b. This result explains why later on the direction of the flow in MGV reverses from up- to down-valley. When it is up-valley, i.e. flowing from RS towards MGV, CIh is about the height of the valley floor at MGV and d ' 500 m (see Fig. 14b). Hence, the upper part of the PCAP above the height of the valley floor at MGV is colder at RS than at MGV (see Fig. 14a), leading to a pressure force between the two sites directed out of the valley. As warm air is advected in the upper part of the PCAP, the capping inversion sharpens (i.e. d decreases) and the temperature in the upper part of the valley atmosphere becomes higher at RS than at MGV (see Fig. 14a and Fig. 14b). This chain of processes leads to the reversal of the pressure force, which is then directed into the valley. As the flow through the tributaries strengthens during the course of the night of P2a, the heat flux through the tributaries increases. From 0200 to 0600 UTC when the mass flux through the tributaries is maximum (see Fig. 10a) and CAPh and CIh decrease continuously with time, ADV is mostly controlled by the heat flux through the tributaries (see Fig. 13a). The analysis of the heat fluxes together with that of Fig. 8, suggest that the upper part of the PCAP is eroded by mixing with the warm air advected through the tributary valleys (mainly MGV after 0200 UTC). Another mechanism that could explain the increase of the potential temperature in the upper part of the PCAP is the downstream displacement of the cold air out of the valley (Lareau and Horel 2015b). However, the comparison of potential temperature profiles at DS and Marnaz does not show evidences of a downstream displacement (not shown). The advection of warm air leads to a continuous decrease of CAPh and CIh during the course of the night, (see Fig. 14b) the latter decreasing to 150 m a.g.l. by 0900 UTC. In WRF-I, the CAP height hardly varies with time (not shown) because of the negligible contribution of the advection from the tributary flow to the heat budget at the valley scale (see Fig. 13d). 5.2.2. Sub-period P2b During P2b the heat fluxes through the tributary valleys are generally weaker than during P2a (see Fig. 13b). The capping inversion subsides down to the surface at 1500UTC, hence leading to a change in the vertical structure of the PCAP. From 1200 to 1500 UTC the channelling of the large-scale flow into the valley through DS led to a substantial inflow into the PCAP, leading to a heat export (cooling) through the top surface of the valley volume. This heat export results from upward vertical motions induced by the convergence of the tributary flows in the upper part of the PCAP and of the channelled flow through DS. Later on, between 1500 UTC to 2000 UTC, turbulent and radiative heat fluxes divergence explain most of the cooling within the PCAP, as the net contribution from advection is close to zero. At 2000 UTC, HDS and HTOP change sign as was the case for the mass fluxes (see Fig. 10a), becoming negative (cooling) and positive (warming), respectively. From this time on and until 1200 UTC, ADV is controlled by the interplay between subsiding motions at the valley top and heat fluxes through MGV and DS, which are modulated by the IGW oscillations of the subsiding motions. 6. Factors controlling the near-surface inversion layer During a PCAP episode, thermally-driven valley flows are one of the main mechanisms favouring the recirculation of air in the near- surface layer (Quimbayo-Duarte et al. 2019). The basin-shape of the Passy Valley constrains horizontal motions in this layer, for which DS is the only gate through which the air mass can flow in and out of the valley. Hence, to investigate the near-surface valley flow, a metric is introduced based on the mass flux through the gate DS (see for instance Whiteman et al. 1996): τ−1 = M∆z100DS [Mass]∆z100 , (4) where [Mass]∆z100 is the mass of the PASSY control volume over the lowest 100 m above the valley floor and M∆z100DS is the mass flux through DS, vertically-integrated over the lowest 100 m above the valley floor. This height was chosen as it approximately corresponds to the height of the nighttime ground-based inversion layer during the persistent stage (see Fig. 14b). The dimension of τ is time, however this parameter is a signed quantity, which sign depends on the direction of the flow through DS. If τ > 0, the flow is out of the valley and τ measures the time required to renew the air mass within the PASSY volume below z = z100 through DS. Conversely, if τ < 0 the flow through DS is into the valley and the near-surface air is nearly stagnant within the valley, assuming that in very stable conditions, vertical motions at z=100 m a.g.l. can be neglected. The parameter τ−1 is displayed in Fig. 15a but, in the following, the behavior of τ is discussed to facilitate its interpretation. τ displays temporal oscillations with a period of approximately 3– 4 hours (see Fig. 15a), which are the signature of the internal gravity waves excited by the flow through the tributary valleys. As pointed out in Sect. 5.1, such oscillations are not present during P2a in WRF-I. The amplitude of the oscillations of τ in WRF-R is fairly large compared to its mean positive value during both P2a and P2b and makes τ to vary between positive (outflow) and negative (inflow) values. During P2a, the average value of τ over a period of oscillation increases with time during the course of the night and the nighttime average is about 7 hours in both WRF-R and WRF-I. The similar trend of τ in WRF-R and WRF-I indicates that the near-surface drainage out of the valley is controlled by local thermally-driven flows but is modulated by the non-local control of the large-scale flow on the flows through the tributaries. The temporal evolution of τ during P2b differs in WRF-R and WRF-I. In WRF-R the average value of τ over a period of oscillation is about 4.5 hours before 2200 UTC and increases to more than 24 hours later on. By contrast, the nighttime average of τ in WRF-I in about 10 hours. This suggests that the non-local control of the large-scale flow is playing a role in the near-surface drainage during P2b in WRF-R. In order to investigate the impact of non-local processes on the thermal structure of the near-surface atmosphere, the tendency (TEND) and advection (ADV) terms of the heat budget (see Sect. 5.2), averaged over the lowest 100 m over the ground surface, are shown for WRF-R and WRF-I in Fig. 15b. The temporal evolution and magnitude of both TEND and ADV during the morning and evening transitions (0700 to 0900 UTC and 1300 to 1500 UTC, respectively) is similar in WRF-R and WRF-I, indicating that local processes control the heat budget of the near-surface atmosphere during these transition periods. During P2a TEND and ADV display a similar trend in WRF-R and WRF-I during nighttime. During P2b, the trends are different: in WRF-R TEND and ADV decrease with time during the night while in WRF-I they increase. Note that as the consequence of the stagnation of the air mass in WRF-R in the late night during P2b, ADV fluctuates about zero. However, the total tendency integrated throughout the night in P2b is very similar between WRF-R and WRF-I (−5.4 K and −5.8 K from sunset to sunrise for WRF-R and WRF-I, respectively), implying that even though the advection contribution varies between the two simulations, the net cooling is hardly affected by the change in the down-valley flow.This article is protected by copyright. All rights reserved. A cc ep te d A rti cl e Figure 15. (a) Timeseries of the inverse of the near-surface (z < 100 m a.g.l.) mass turnover time τ associated with the mass flux through the downstream (DS) gate (see Fig. 1b) for the period from 10 February 2015 at 1200 UTC to 12 February 2015 at 1200 UTC in WRF-R and WRF-I. (b) Timeseries of the density-weighted total potential temperature tendency (TEND) and advection contribution (ADV) to the total tendency, averaged over the PASSY volume over the lowest 100 m above the ground surface, for the period from 10 February 2015 at 1200 UTC to 12 February 2015 at 1200 UTC in WRF-R and WRF-I. The nighttime periods (from sunset to sunrise) are displayed as grey shading. Figure 16. Timeseries of the pressure difference (black lines) between the RS and MRNZ sites (seeFig. 1b) at z= 600 m a.s.l. (∆P600) and at z= zm = 1560 m a.s.l. (∆P1560), in (a) WRF-R and (b) WRF-I. Each panel also display the difference in potential temperature (blue line) between RS and MRNZ sites, vertically averaged in the height range 600 < z < zm = 1560m a.s.l. (∆θRS-MRNZ). The nighttime periods (from sunset to sunrise) are displayed as grey shading. This difference between WRF-R and WRF-I during P2b is a result of the evolution of the large-scale flow. The strengthening of the south-southeasterly large-scale flow during the course of the night favours its channelling through MGV in the upper part of the PCAP. The large PCAP stability prevents the channeled flow to penetrate down at the valley bottom and so to enhance the down-valley flow. The channelled flow transports warmer air in the upper part of the PCAP, so reducing the horizontal temperature difference between the valley centre and outside the valley towards MRNZ (see Fig. 13a and Fig. 16a). This results in a reduction in the near-surface pressure difference driving the down-valley flow between RS and MRNZ between 2100 and 0000 UTC (see Fig. 16a), leading to the stagnation of the near- surface air within the valley. This non-local effect is not present in WRF-I, which accounts for the near-surface pressure difference to continue to drive a down-valley flow throughout the night (see Fig. 16b). 7. Summary and conclusions The purpose of the present work is to examine local and non- local controls on a persistent cold-air pool (PCAP) that formed in the Passy Valley in February 2015 using numerical model simulations. The effect of the non-local controls on the PCAP was determined by comparing the real case simulation with a semi- idealised simulation with initial quiescent synoptic conditions. The Passy Valley is a region of Alpine terrain with steep slopes and three major tributary valleys, Mege`ve, Saint-Gervais- les-Bains and Chamonix (see Sect. 2.1). The numerical model simulations were performed using the Weather Research and Forecasting model at a horizontal mesh size of 111 m, carefully initialised (see Sect. 2 for details). The simulation results are found to realistically capture features of the persistent cold-air pool by means of comparison with field observations. The model realistically simulates the magnitude and temporal evolution of the valley heat deficit, as well as the vertical structure of the potential temperature, when compared to the available observations. However, it overestimates the near-surface minimum temperature at nighttime, which leads to an underestimation of the magnitude of the temperature inversion over the lowest 100m. The cold-air pool formed in the evening on 9 February 2015 as a result of the radiative cooling at the valley floor and advection of warm air above the Passy valley at the mean height of the terrain surrounding the valley, which sets the height of the cold- air pool during its formation stage. The advection of warm air was associated with the passage of an upper level ridge over the region. The effects of the large-scale flow on the cold-air pool during its persistent stage, from 10 February 2015 at 1200 UTC to 12 February 2015 at 1200 UTC, was investigated at the valley scale and in the near-surface layer above the valley floor (lowest 100 m). The analysis of the numerical simulations showed that the mass fluxes at the valley scale through the tributaries are about one order of magnitude larger than in the semi-idealised simulation, which demonstrates the impact of the large scale flow on the flow through the tributary valleys. The flow through Chamonix and Saint-Gervais-les-Bains valleys can supply up to 50% of the air mass that drained out of the valley during one of the nights of the persistent stage of the CAP. As a consequence, the heat budget at the valley scale differs between the real andThis article is protected by copyright. All rights reserved. A cc ep te d A rti cl e idealised simulations: this budget is controlled primarily by the flows through the tributary valleys and the vertical advection at the valley top in the real case. This combination of horizontal and vertical advection modified the height of the PCAP. This led to a change in the direction of the flow in the Mege`ve tributary valley, which substantially modified the dynamics and thermal structure of the Passy Valley atmosphere. The results thus showed that local thermally driven flows alone cannot explain the day-to- day variability of the CAP height. Conversely, the near-surface atmosphere was found to respond differently depending on the position of the upper level ridge. When the upper level ridge passed over the Passy Valley (between 10 and 11 February 2015), the mass and heat budgets computed for a near-surface atmospheric volume were found to be comparable between the real and semi-idealised simulations. This implies that the near-surface atmosphere above the valley floor was nearly decoupled from the upper part of the cold-air pool and controlled by and large by local processes, although the flow was modulated by internal gravity waves generated by the detachment of the flow from one of the tributary valleys. As the upper level ridge moved away from the Passy Valley (between 11 and 12 February 2015), the large-scale flow strengthened, favouring channelling through the Mege`ve tributary valley towards Passy. This flow transported warmer air (in terms of potential temperature) in the upper part of the PCAP, reducing the thermal gradient in the down-valley direction, while being decoupled from the near-surface atmosphere because of the large PCAP stability. This leads to the reduction of the surface pressure difference driving the down-valley flow, eventually leading to stagnant air over the valley floor. This mechanism is not present in the semi-idealised simulation, in which a down-valley flow was found to flow out of the Passy Valley throughout the night. In summary, the main findings of the present work, which are of relevance for other deep Alpine valleys under similar wintertime anticyclonic conditions, are: • The large-scale flow controls the magnitude of the mass fluxes through the tributary valleys and their relative contribution to the total flow out of the Passy Valley. The flows from the tributaries can contribute more than 50% to this outflow during the persistent stage of the episode. • The direction of the flow through one of the tributaries is found to be determined by the height of the PCAP with respect to that of the tributary above the valley floor. This result highlights the coupling between the large-scale flow, the tributary flows and the thermal structure of the valley atmosphere during a PCAP event. • The tributary and large-scale flows modify the heat budget of the valley atmosphere during the persistent stage of the episode, when compared to the heat budget for quiescent synoptic conditions. At the valley scale, the advection contribution is found to be the main control on the total potential temperature tendency. • Hence, the effect of the non-local controls is key for determining the change in the thermal structure of the CAP from one day to the following and so the change in the heat deficit. It follows that local thermally-driven flows alone are not able to modify the thermal structure of the PCAP on the time scale of this episode. • In the present case, the near-surface atmosphere remains almost decoupled from the flow aloft, and two mechanisms are identified. When the upper level ridge passes over the region, the near-surface atmosphere is subject to usual down-valley flows during nighttime; when the ridge moves away from the region, the near-surface atmosphere become nearly stagnant. This is due to the warm air advection in the upper part of the CAP by the large-scale flow. The present work calls for a better characterisation of the controls of tributary flows and of varying large-scale flows on cold-air pooling. This would require an approach based on a climatology of the area to identify the prevalence of weather patterns. Future field campaigns should consider instrumenting the main tributary valleys and collecting long-term datasets to sample an extended range of synoptic forcings. Further observations are also required in order to evaluate the role of internal gravity waves in the modulation of the heat and mass budgets of the valley atmosphere. Finally, future modelling studies should also consider the role of fog, and more generally of microphysics processes, in the evolution of a PCAP, as these were not considered in the current work. Acknowledgements This work has been supported by a grant from LabEx Osug@2020 (Investissements d’avenir – ANR10LABX56). Numerical model simulations were performed using the UK national supercomputing facilities (ARCHER), accessed through the UK National Centre for Atmospheric Science (NCAS). The Passy-2015 field experiment was supported by ADEME through the French national programme LEFE/INSU and by METEO- FRANCE. We thank the cities of Passy and Sallanches for their kind support. The field experiment was led by CNRM while the LEGI laboratory was the principal investigator of the LEFE/INSU project. The rawinsonde data used in Fig. 3 are managed by SEDOO at Observatoire Midi-Pyre´ne´es. We thank the people involved in the launching of the rawinsonde and data acquisition therefrom. References Arduini G, Chemel C, Staquet C. 2017. Energetics of deep alpine valleys in pooling and draining configurations. J. Atmos. Sci. 74: 2105–2124. Arduini G, Staquet C, Chemel C. 2016. Interactions between the nighttime valley wind system and a developing cold-air pool. Boundary- Layer Meteorol. 161: 49–72. Billings BJ, Grubisˇic V, Borys RD. 2006. Maintenance of a mountain valley cold pool: a numerical study. Mon. Weather. Rev. 134: 2266–2278. Bodine D, Klein PM, Arms SC, Shapiro A. 2009. Variability of surface air temperature over gently sloped terrain. J. Appl. Meteorol. Climatol. 48: 1117–1141. Burns P, Chemel C. 2015. Interactions between downslope flows and a developing cold-air pool. Boundary-Layer Meteorol. 154: 57–80. Chemel C, Arduini G, Staquet C, Largeron Y, Legain D, Tzanos D, Paci A. 2016. Valley heat deficit as a bulk measure of wintertime particulate air pollution in the Arve River Valley. Atmos. Environ. 128: 208–216. Chen F, Dudhia J. 2001. Coupling an advanced land-surface/hydrology model with the Penn State/NCAR MM5 modeling system. Part I: model implementation and sensitivity. Mon. Weather. Rev. 129: 569–585. Chow FK, Weigel AP, Street RL, Rotach MW, Xue M. 2006. High-resolution Large-Eddy Simulations of flow in a steep alpine valley. Part I: Methodol- ogy, verification, and sensitivity experiments. J. Appl. Meteorol. Climatol. 45(1): 63–86. Clements CB, Archuleta J, Gudiksen PH. 1989. Experimental design of the 1984 ASCOT field study. J. Appl. Meteorol. 28: 405–413. Clements CB, Whiteman CD, Horel JD. 2003. Cold-air-pool structure and evolution in a mountain basin: Peter Sinks, Utah. J. Appl. Meteorol. 42: 752–768. Coulter RL, Martin TJ, Porch WM. 1991. A comparison of nocturnal drainage flow in three tributaries. J. Appl. Meteorol. 30: 157–169. Coulter RL, Orgill M, Porch W. 1989. Tributary fluxes into Brush Creek Valley. J. Appl. Meteorol. 28: 555–568. Farr TG, Rosen PA, Caro E, Crippen R, Duren R, Hensley S, Kobrick M, Paller M, Rodriguez E, Roth L, Seal D, Shaffer S, Shimada J, Umland J, Werner M, Oskin M, Burbank D, Alsdorf D. 2007. The Shuttle Radar Topography Mission. Rev. Geophys. 45, doi:10.1029/2005RG000183. Flamant C, Drobinski P, Furger M, Chimani B, Tschannett S, Steinacker R, Protat A, Richner H, Gubser S, Ha¨berli C. 2006. Fo¨hn/cold-pool interactions in the Rhine valley during MAP IOP 15. Q. J. R. Meteorol. Soc. 132: 3035–3058.This article is protected by copyright. All rights reserved. A cc ep te d A rti cl e Hall DK, Riggs GA, Salomonson VV. 2006. MODIS/Terra Snow Cover 5-Min L2 Swath 500m, Version 5. Technical report, Boulder, CO, USA. NASA National Snow and Ice Data Center Distributed Active Archive Center, doi: 10.5067/ACYTYZB9BEOS. Hong SY. 2010. A new stable boundary-layer mixing scheme and its impact on the simulated East Asian summer monsoon. Q. J. R. Meteorol. Soc. 136: 1481–1496. Jime´nez PA, Dudhia J, Gonza´lez-Rouco JF, Navarro J, Monta´vez JP, Garcı´a- Bustamante E. 2012. A revised scheme for the WRF surface layer formulation. Mon. Weather. Rev. 140(3): 898–918. Kiefer MT, Zhong S. 2011. An idealized modeling study of nocturnal cooling processes inside a small enclosed basin. J. Geophys. Res. 116. D20127, doi:10.1029/2011JD016119. Lareau NP, Crosman E, Whiteman CD, Horel JD, Hoch SW, Brown WOJ, Horst T. 2013. The Persistent Cold-Air Pool Study. Bull. Am. Meteorol. Soc. 94: 51–63. Lareau NP, Horel J. 2015a. Turbulent erosion of persistent cold-air pools: numerical simulations. J. Atmos. Sci. 72: 1409–1427. Lareau NP, Horel JD. 2015b. Dynamically induced displacements of a persistent cold-air pool. Boundary-Layer Meteorol. 154(2): 291–316. Largeron Y, Staquet C. 2016a. The atmospheric boundary layer during wintertime persistent inversions in the Grenoble valleys. Front Earth Sci. doi:10.3389/feart.2016.00070. Largeron Y, Staquet C. 2016b. Persistent inversion dynamics and wintertime PM10 air pollution in Alpine valleys. Atmos. Environ. 135: 92–108. Largeron Y, Staquet C, Chemel C. 2013. Characterization of oscillatory motions in the stable atmosphere of a deep valley. Boundary- Layer Meteorol. 148: 439–454. Lee T, Fukumori I, Tang B. 2004. Temperature advection: internal versus external processes. J. Phys. Oceanogr. 34: 1936–1944. Livneh B, Xia Y, Mitchell KE, Ek MB, Lettenmaier DP. 2010. Noah LSM snow model diagnostics and enhancements. J. Hydrol. 11: 721–738. Lu W, Zhong S. 2014. A numerical study of a persistent cold air pool episode in the Salt Lake Valley, Utah. J. Geophys. Res. 119: 1733–1752. McKee TB, O’Neal RD. 1989. The role of valley geometry and energy budget in the formation of nocturnal valley winds. J. Appl. Meteorol. 28: 445–456. Meløysund V, Leira B, Høiseth KV, Lisø KR. 2007. Review – Predicting snow density using meteorological data. Meteorol. Appl. 14: 413–423. Mlawer EJ, Taubman SJ, Brown PD, Iacono MJ, Clough SA. 1997. Radiative transfer for inhomogeneous atmospheres: RRTM, a validated correlated-k model for the longwave. J. Geophys. Res. 102: 663–682. Morrison H, Curry JA, Khvorostyanov VI. 2005. A new double-moment microphysics parameterization for application in cloud and climate models. Part I: Description. J. Atmos. Sci. 62: 1665–1677. Nakanishi M. 2000. Large-Eddy Simulation of radiation fog. Boundary- Layer Meteorol. 94(3): 461–493. Neemann EM, Crosman ET, Horel JD, Avey L. 2015. Simulations of a cold-air pool associated with elevated wintertime ozone in the Uintah basin, Utah. Atmos. Chem. Phys. 15: 135–151. O’Steen LB. 2000. Numerical simulation of nocturnal drainage flows in idealized valley-tributary system. J. Appl. Meteorol. 39: 1845–1860. Paci A, Staquet C, Allard J, Barral H, Canut G, Cohard J, Jaffrezo J, Martinet P, Sabatier T, Troude F, et al. 2016. La campagne passy-2015: dynamique atmosphe´rique et qualite´ de lair dans la valle´e de larve. Poll. Atmos. : 231– 232. Page`s M, Pepin N, Miro´ JR. 2017. Measurement and modelling of temperature cold pools in the Cerdanya valley (Pyrenees), Spain. Meteorol. Appl. 24(2): 290–302. Porch WM, Fritz RB, Coulter RL, Gudiksen PH. 1989. Tributary, valley and sidewall air flow interactions in a deep valley. J. Appl. Meteorol. 28: 578– 589. Quimbayo-Duarte J, Staquet C, Chemel C, Arduini G. 2019. Dispersion of tracers in the stable atmosphere of a valley opening onto a plain. Boundary- Layer Meteorol. 172: 291–315. Reeves HD, Stensrud DJ. 2009. Synoptic-scale flow and valley cold pool evolution in the Western United States. Weath. Forecasting 24: 1625–1643. Robinson DA, Kukla G. 1985. Maximum surface albedo of seasonally snow- covered lands in the northern hemisphere. J. Climate Appl. Meteor. 24: 402– 411. Schmidli J, Poulos GS, Daniels MH, Chow FK. 2009. External influences on nocturnal thermally driven flows in a deep valley. J. Appl. Meteorol. Clima- tol. 48: 3–23. Schmidli J, Rotunno R. 2010. Mechanisms of along-valley winds and heat exchange over mountainous terrain. J. Atmos. Sci. 67(9): 3033–3047. Sheridan PF. 2019. Synoptic-flow interaction with valley cold-air pools and effects on cold-air pool persistence: Influence of valley size and atmospheric stability. Q. J. R. Meteorol. Soc. 145: 1636–1659. Silcox GD, Kelly KE, Crosman ET, Whiteman CD, Allen BL. 2012. Wintertime PM25˙ concentrations during persistent, multi-day cold-air pools in a mountain valley. Atmos. Environ. 46: 17–24. Skamarock WC, Klemp JB, Dudhia J, Gill DO, Barker DM, Duda MG, Huang XY, Wang W, Powers JG. 2008. A description of the Advanced Research WRF Version 3. NCAR Technical Note NCAR/TN-475+STR, NCAR, Boulder, CO, USA. 125 pp. Staquet C, Paci A, 24 co-authors. 2015. The Passy project: objectives, underlying scientific questions and preliminary numerical modelling of the Passy Alpine valley. In: Proc. of the 33rd International Conference on Alpine Meteorology, Innsbruck, Austria. Tomasi E, Giovannini L, Zardi D, De Franceschi M. 2017. Optimization of Noah and Noah MP WRF land surface schemes in snow-melting conditions over complex terrain. Mon. Weather. Rev. 145: 4727–4745. Vosper SB, Brown AR. 2008. Numerical simulations of sheltering in valleys: the formation of nighttime cold-air pools. Boundary-Layer Meteorol. 127(3): 429–448. Vosper SB, Hughes JK, Lock AP, Sheridan PF, Ross AN, Jemmett- Smith B, Brown AR. 2014. Cold-pool formation in a narrow valley. Q. J. R. Meteorol. Soc. 140: 699–714. Vrhovec T. 1991. A cold air lake formation in a basin – a simulation with a mesoscale numerical model. Meteorol. Atmos. Phys. 46: 91–99. Vrhovec T, Hrabar A. 1996. Numerical simulations of dissipation of dry temperature inversions in basins. Geofizika 13: 81–96. Wei L, Zhaoxia P, Shigong W. 2013. Numerical simulation of the life cycle of a persistent wintertime inversion over Salt Lake City. Boundary- Layer Meteorol. 148: 399–418. Whiteman CD. 1982. Breakup of temperature inversions in deep mountain valleys: Part I. Observations. J. Appl. Meteorol. 21: 270–289. Whiteman CD. 1990. Observations of thermally developed wind systems in mountainous terrain. In: Atmospheric processes over complex terrain, Meteorological Monographs, vol. 23, no. 45, Blumen W (ed), ch. 2, American Meteorological Society, 45 Beacon St., Boston, MA, USA, pp. 5–42. Whiteman CD, Bian X, Zhong S. 1999. Wintertime evolution of the temperature inversion in the Colorado Plateau Basin. J. Appl. Meteorol. 38: 1103–1117. Whiteman CD, McKee TB, Doran JC. 1996. Boundary layer evolution within a canyonland basin. Part I: Mass, heat, and moisture budgets from observations. J. Appl. Meteorol. 35: 2145–2161. Za¨ngl G. 2005a. Dynamical aspects of wintertime cold-air pools in an alpine valley system. Mon. Weather. Rev. 133: 2721–2740. Za¨ngl G. 2005b. Formation of extreme cold-air pools in elevated sinkholes: an idealized numerical process study. Mon. Weather. Rev. 133: 925–941. Za¨ngl G. 2005c. Wintertime cold-air pools in the Bavarian Danube Valley basin: data analysis and idealized numerical simulations. J. Appl. Meteorol. 44: 1959–1971. Zhong S, Bian X, Whiteman CD. 2003. Time scale for cold-air pool breakup by turbulent erosion. Meteorol. Z. 12: 229–233. Zhong S, Whiteman CD, Bian X, Shaw WJ, M HJ. 2001. Meteorological processes affecting the evolution of a wintertime cold air pool in the Columbia Basin. Mon. Weather. Rev. 129: 2600–2613. This article is protected by copyright. All rights reserved. A cc ep te d A rti cl e