Polar vortex concentricity as a controlling factor in Arctic denitrification

Size: px
Start display at page:

Download "Polar vortex concentricity as a controlling factor in Arctic denitrification"

Transcription

1 JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 107, NO. D22, 4663, doi: /2002jd002102, 2002 Polar vortex concentricity as a controlling factor in Arctic denitrification G. W. Mann, S. Davies, K. S. Carslaw, and M. P. Chipperfield Institute for Atmospheric Science, School of the Environment, University of Leeds, Leeds, UK J. Kettleborough British Atmospheric Data Centre, Rutherford Appleton Laboratory, Didcot, UK Received 16 January 2002; revised 21 May 2002; accepted 7 June 2002; published 30 November [1] Recent in situ observations in the Arctic stratosphere have detected nitric acid containing particles with sizes up to 10-mm radius and number concentrations between 10 5 and 10 3 cm 3. Here we quantify the effect of these particles on Arctic denitrification by using a new three-dimensional (3-D) model which can couple particle growth and sedimentation with the full dynamics of the Arctic polar vortex. We show that the very long growth times of large nitric acid trihydrate (NAT) particles lead to a highly nonlinear dependence of Arctic denitrification on the growth and evaporation cycles of individual particles, thus making denitrification dependent on the precise meteorological conditions in a given winter. Using 3-D wind and temperature fields from December 1999, we identify a period that was optimum for denitrification, in which the cold pool and vortex were nearly concentric and in which a large proportion of the particles were able to sediment over about 8 days through the full depth of the cold pool without evaporating. We then show that small departures from concentric conditions can lead to substantial reductions in denitrification. A case is presented in which denitrification was completely shut off even with over half of the cold pool area remaining within the vortex. Under the same conditions, a model in which the particles were assumed to be in continuous equilibrium with the gas phase caused extensive denitrification. Our results show that low Arctic vortex temperatures in themselves are unlikely to be a reliable indicator of potential denitrification if large NAT particles are involved. INDEX TERMS: 0305 Atmospheric Composition and Structure: Aerosols and particles (0345, 4801); 0340 Atmospheric Composition and Structure: Middle atmosphere composition and chemistry; 3334 Meteorology and Atmospheric Dynamics: Middle atmosphere dynamics (0341, 0342); 3337 Meteorology and Atmospheric Dynamics: Numerical modeling and data assimilation; KEYWORDS: polar stratospheric clouds, NAT rocks, denitrification, stratosphere, Arctic, vortex Citation: Mann, G. W., S. Davies, K. S. Carslaw, M. P. Chipperfield, and J. Kettleborough, Polar vortex concentricity as a controlling factor in Arctic denitrification, J. Geophys. Res., 107(D22), 4663, doi: /2002jd002102, Introduction Copyright 2002 by the American Geophysical Union /02/2002JD [2] Denitrification is the irreversible removal of HNO 3 from the polar stratosphere due to the gravitational sedimentation of HNO 3 -containing particles that form in the polar winter. Under most Arctic and Antarctic conditions, the reductions in gas phase HNO 3 concentrations lead to increased ozone loss rates in springtime. This increased loss rate results from increased active chlorine concentrations, which would otherwise be buffered by reactive nitrogen compounds released from HNO 3 photolysis or reaction between OH and HNO 3 (see, e.g., Solomon [1999]). Several chemical model simulations show that the additional cumulative ozone loss on a single level caused by denitrification of the Arctic stratosphere can be as much as 30% [Chipperfield and Pyle, 1998; Waibel et al., 1999; Tabazadeh et al., 2000]. [3] Denitrification is a ubiquitous feature of the Antarctic polar winter stratosphere, where temperatures are persistently cold. Denitrification has also been observed in the Arctic in the cold winters 1988/1989, 1994/1995, 1995/1996, 1996/1997 and 1999/2000 [Fahey et al., 1989; Hubler et al., 1990; Rex et al., 1997; Hintsa et al., 1998; Waibel et al., 1999; Kondo et al., 1999, 2000; Dessler et al., 1999; Santee et al., 1999, 2000; Popp et al., 2001; Kleinböhlet al., 2002]. [4] An important parameter in any model of denitrification is the number density and size of the sedimenting particles. Arctic observations during winter 1999/2000 detected very low number concentrations ( cm 3 ) of large (up to 10-mm radius) nitric acid containing particles [Fahey et al., 2001; Northway et al., 2002]. These observations have led to a number of model studies aimed at understanding the evolution of such particles [Carslaw et AAC 13-1

2 AAC 13-2 MANN ET AL.: VORTEX CONCENTRICITY AND ARCTIC DENITRIFICATION al., 2002] as well as their effect on denitrification [Davies et al., 2002; Jensen et al., 2002]. The model study of Tabazadeh et al. [2001], while not motivated directly by these observations, also explored the effect of large nitric acid trihydrate (NAT) particles on denitrification. [5] At present, it is not known how common such large particles are in the Arctic stratosphere because, aside from the NO y instrument used during the 1999/2000 winter, the observing systems needed for their detection have not been deployed. However, their existence has important implications for the way that NAT particles and denitrification are treated in models. In particular, the particle sizes require continuous growth periods for individual particles of typically 5 8 days [Fahey et al., 2001; Carslaw et al., 2002]. However, the typical meteorology of the Arctic stratosphere is such that sedimenting particles are likely to be advected into regions above the NAT temperature more often than once in 5 8 days, implying that large NAT particles may not be able to form in all winters. At the low number densities of hydrate particles observed, the NAT particles will not be in equilibrium with the gas phase HNO 3 [Carslaw et al., 2002]. Thus models must account for the time-dependent growth and evaporation of sedimenting particles throughout the vortex. At present, all three-dimensional (3-D) chemical models assume polar stratospheric clouds (PSC) to be at thermodynamic equilibrium (see, for example, Chipperfield [1999], Considine et al. [2000], and Davies et al. [2002]). [6] Jensen et al. [2002] used a 1-D column model combined with artificial temperature histories to explore the effect on denitrification of the length of the cooling cycle, together with other parameters such as particle number density, and uptake of HNO 3 by liquid aerosols. As was expected, they found that denitrification was increased if particles were allowed to grow for longer periods to larger sizes, at least when number concentrations were low. [7] Tabazadeh et al. [2000] showed that the magnitude of denitrification could be linked to the length of time that air parcels spend below the NAT temperature. They related the observed denitrification to the length of time that air parcels on the 450 K potential temperature surface spent below the NAT temperature. However, an analysis of isentropic air parcels is inappropriate when low number densities of growing particles are involved, since the particles quickly sediment to lower atmospheric layers. The rapid sedimentation of particles (a 6-mm radius particle has a fall speed of 0.9 km per day compared with 0.1 km per day for a 2-mm radius particle) means that isentropic trajectories are valid only for periods shorter than 1 or 2 days. Rather, it is the temperature histories experienced by the particles themselves that are important, and they differ markedly from isentropic trajectories. It is such individual particle trajectories which we examine here as a controlling factor in denitrification. Such a particle analysis is possible only with a model that tracks individual particles in three dimensions [Carslaw et al., 2002]. [8] The idealized 1-D model study of Jensen et al. [2002] was a useful exploration of many of the factors that control denitrification. However, the simulations were restricted to single columns of air in the vortex and hence do not represent the full complexity of coupled particle and air motion in three dimensions. The 3-D model study of Carslaw et al. [2002] focused on the evolution of the large NAT particles and demonstrated that the change in particle characteristics observed in the period January to March 2000 [Northway et al., 2002] could be explained in terms of changes in the 3-D temperature and flow fields. However, they did not calculate denitrification. [9] Here we apply the 3-D microphysical model of Carslaw et al. [2002] to understand how the variable meteorology of the Arctic stratosphere controls denitrification by large NAT particles. In section 2 we describe the model structure and the general setup of the model runs. In section 3 we describe the motivation for and design of the numerical experiments. In section 4 the denitrification results are presented and interpreted in terms of the behavior of the NAT particles in the model. 2. Model Structure [10] The model is described in detail by Carslaw et al. [2002]. Briefly, the model simulates the time-dependent growth and evaporation of several thousand individual nitric acid hydrate particles, with their 3-D motion calculated by using isentropic trajectories combined with vertical motion due to gravitational sedimentation, which depends on particle size. Particle growth by diffusive transfer of HNO 3 removes HNO 3 from the gas phase in each model time step. Gas phase HNO 3 is then calculated (with chemistry switched off) on a 3-D grid and advected each time step using the off-line global Eulerian chemical transport model SLIMCAT [Chipperfield et al., 1996]. For all runs in the present study we have used a SLIMCAT grid with 36 vertical levels between 350 K and 2720 K, giving a resolution of K in the lower stratosphere. [11] The equation for particle growth is where G ¼ D HNO 3 M NAT 2r NAT RT r 2 ¼ r 2 0 þ Gt; ð1þ p HNO3 p NAT HNO 3 : ð2þ D HNO3 is the diffusion coefficient of HNO 3 molecules in air, modified to account for mass transfer noncontinuum effects for particles with size similar to the mean free path. M NAT is the molar mass of NAT and r NAT is the NAT crystal mass NAT density, while p HNO3 and p HNO3 are the ambient partial pressure of HNO 3 and the vapor pressure of HNO 3 over NAT, respectively. Particles are assumed to be in the Stokes drag regime with the sedimentation velocity given by where dz dt ¼ Sr2 ; S ¼ 2gr NAT C C 9h is the sedimentation factor. C C is the Cunningham slip flow correction factor [Pruppacher and Klett, 1997] and h is the viscosity of air. ð3þ ð4þ

3 MANN ET AL.: VORTEX CONCENTRICITY AND ARCTIC DENITRIFICATION AAC 13-3 [12] For very small particles with negligible sedimentation speed, the combined Lagrangian advection of the particulate HNO 3 and the Eulerian advection of the gas phase HNO 3 conserves total HNO 3 and is equivalent to the straightforward isentropic advection of a tracer in an Eulerian model. In the case of large particles that sediment, the separate advection of gas and particulate HNO 3 leads to 3-D fields of denitrification and renitrification. At each time step, the gas phase HNO 3 is used to calculate subsequent particle growth, so that localized gas phase HNO 3 removal by the hydrate particles feeds back into a reduced particle growth rate. By simulating the motion of the gas phase HNO 3 and particles simultaneously in 3-D space, the model does not make any simplifying assumptions about particle temperature histories or available HNO 3, which are typical deficiencies of column models. [13] In this study we take account only of the HNO 3 partitioning to NAT particles and neglect the uptake into liquid HNO 3 /H 2 SO 4 /H 2 O droplets, which occurs at temperatures typically 4 K below the NAT condensation temperature [Carslaw et al., 1995]. This simplification is justified in the present study since temperatures remain marginally above the temperature at which uptake into droplets occurs. [14] Nucleation of new NAT particles is treated in a straightforward manner. At each time step (here 30 min), new NAT particles are created with a radius of 0.1 mm inall NAT-supersaturated grid boxes of a cosine-weighted horizontal grid and at randomly chosen altitudes. The average nucleation rate (particles per unit air volume and time) is uniform throughout the NAT-supersaturated region. Random altitude levels of nucleation ensure that coherent structures in the final particle size distributions at any one level are smoothed out. Other nucleation mechanisms can also be used [Carslaw et al., 2002]. We have chosen an average nucleation rate corresponding to particle cm 3 s 1, which produces a total number concentration of large NAT particles (radius of >2 mm) equal to cm 3 along the ER-2 flight track on 20 January 2000 [Carslaw et al., 2002], which compares well with the observations of Fahey et al. [2001]. The model transports approximately 20,000 particles at any one time; thus one model particle represents a much higher number of real particles. [15] So long as the nucleation mechanism for very low number concentrations of NAT particles remains unknown, our spatially uniform formation mechanism is the most straightforward method of producing vortex-scale populations of large NAT particles with sizes and number concentrations in good agreement with observations [Fahey et al., 2001; Northway et al., 2002; Carslaw et al., 2002]. Although details of the results may differ with a different nucleation mechanism, our conclusions are unlikely to be sensitive to the precise nucleation mechanism, particularly as our present study focuses on the evolution of the particles in the several days after spatially uniform populations have been generated. 3. Description of the Simulations 3.1. Wind and Temperature Fields [16] The aim of this study is to understand how the characteristic meteorology of the Arctic stratosphere controls denitrification. In the Antarctic polar stratosphere the wind and temperature fields are normally concentric about the pole. In such a state, the region of low temperatures in which PSCs can form (the cold pool) is normally contained within the vortex. In contrast, the Arctic wind and temperature fields are often significantly offset (a strongly baroclinic state) [Pawson et al., 1995]. This characteristic feature of the Arctic stratosphere is likely to exert a strong influence on denitrification by low number densities of slowly growing NAT particles. When the cold pool and the vortex are concentric, NAT particles will tend to remain inside the cold pool as they are advected by the wind, resulting in large particles, but when the cold pool and the vortex are offset, NAT particles are more likely to be advected out of the cold pool and into warmer air before they have sufficient time to grow large enough to sediment out. It is this link between the meteorology of the Arctic stratosphere, the growth period of individual particles, and hence the amount of denitrification that we investigate here in a carefully controlled manner. [17] Pawson et al. [1995] analyzed 30-hPa and 50-hPa geopotential height and temperature fields from winters 1965/1966 to 1993/1994 and described several typical configurations of the Arctic stratospheric flow. These configurations range from near-concentric cases when the cold pool center is colocated with the vortex center to disturbed, nonconcentric cases when the cold pool is displaced considerably from the center of the vortex. The analysis of Pawson and Naujokat [1999] clearly shows that such nonconcentric vortex states do occur in the Arctic stratosphere sufficiently frequently and for sufficient periods of time to influence particle temperature trajectories and, hence, denitrification. [18] In order to isolate the influence of variations in Arctic flow on denitrification, we have used a very controlled set of meteorological conditions and eliminated variations in other factors that influence denitrification. Idealized model runs simulating a development over 10 days were carried out by using fixed wind and temperature fields from the European Centre for Medium-Range Weather Forecasts (ECMWF) as the model forcing. Simulations of 10 days are long enough for particles nucleated in the uppermost level of the cold pool to sediment to the bottom of the stratosphere. In our baseline run we used ECMWF-analyzed wind and temperature fields from 23 December 1999 chosen as a good example of where the cold pool and vortex are nearly concentric or Antarctic-like. We will refer to this configuration of the Arctic stratosphere as the concentric state. Nonconcentric states were produced by artificially shifting the cold pool center away from the vortex center but retaining an identical cold pool shape and area. By doing this, cycles of high and low temperatures experienced by the NAT particles could be influenced without affecting the total area of PSC formation. The shift in the cold pool was achieved by using solid body rotations of between 2.5 and 20, the axis of rotation being the line connecting latitude/longitude points [0, 0 ] and [0, 180 ] on opposite sides of the sphere. The 0 rotation represents a real near-concentric vortex state, while the 20 rotation of the temperature field produces a temperature field representing an idealized cold strongly baroclinic state of the Arctic stratosphere.

4 AAC 13-4 MANN ET AL.: VORTEX CONCENTRICITY AND ARCTIC DENITRIFICATION Figure 1. Potential vorticity and temperature fields used in the 3-D simulations. Fields are shown for the 585 K potential temperature level because it is the height of the temperature minimum. The angles of each solid body rotation of the temperature field are indicated on each plot. The thick black line shows the vortex edge, as defined by a modified potential vorticity of 40 PV units (i.e., km 2 kg 1 s 1 ). The thick white line is the isopleth of T NAT. Within this region the air is supersaturated with respect to NAT. The 0 case is referred to in the text as the concentric state. Initial fields of HNO 3 and H 2 O were assumed spatially uniform at 8 ppbv and 5 ppmv. [19] Figure 1 shows the temperature fields for 0, 10, 15, and 20 rotations on the 585 K potential temperature surface. This height was chosen because it was the height of the temperature minimum for this ECMWF analysis. For reference, the vortex edge is marked on each plot as a thick black line. Throughout this paper, we define the vortex edge as where the modified potential vorticity (MPV) is 40 PV units (i.e., km 2 kg 1 s 1 ). Modified potential vorticity, is defined as Ertel s potential vorticity (EPV) multiplied by the scaling factor ðq=q 0 Þ 9 2 ; where q0 is a reference potential temperature (taken as 475 K), hence removing much of the density dependence of EPV [Lait, 1994]. We have taken the value of 40 PV units as it is seen to correspond well with the maximum PV gradient, the formal definition of the vortex edge. The white line marks the saturation temperature for NAT, T NAT, calculated by using the expression from Hanson and Mauersberger [1988]. The centroids of the vortex and NAT region are shown for this q surface as black and white squares, respectively. For the 0 run, the entire NAT region is inside the vortex and the centroids are approximately colocated, while for the 20 run, the centroid of the NAT region lies approximately on the vortex edge, and only around half of the NAT region is inside the polar vortex on this q surface. We use the separation of the vortex and NAT region centroids as a measure of vortex concentricity. The centroid separations are normalized to the effective vortex radius at each level (R eff =(A/p) 1/2, where A is the vortex area). A normalized centroid separation equal to 0 implies that the vortex and cold pool centroids are colocated (a highly concentric vortex), and a normalized separation equal to 1 implies that the cold pool centroid sits on the vortex edge (a nonconcentric vortex). [20] Our method of shifting the temperature field means that the wind, temperature, and pressure fields are no longer in thermal wind balance in the rotated runs. However, our intention is to carefully control the cooling/heating cycles of the particles being advected around the vortex, and the nonphysical nature of the meteorological fields that we use are of no consequence in this respect. Fixing the wind, pressure, and temperature fields for 10 days is also somewhat unrealistic. In reality, the cold pool will, to some extent, move around; thus the effect of nonconcentricity on the temperature histories of particles may be smaller in

5 MANN ET AL.: VORTEX CONCENTRICITY AND ARCTIC DENITRIFICATION AAC 13-5 Figure 2. Relative location of the NAT region and vortex as a function of altitude and rotation of the temperature field, as indicated by the separation of their centroids. reality. However, Pawson et al. [1995] show time series of the magnitude of the geostrophic wind velocity from winters 1984/1985 to 1993/1994 which show numerous periods (e.g., February 1986, January 1987, and February 1993) where the cold center is at the edge of the vortex (the geostrophic wind is high and the flow is strongly baroclinic) and the cold center has T < 195 K for long continuous periods. Consequently, these artificially produced vortex/ cold pool states do represent realistic cases of stable polar vortices of differing baroclinicity. [21] Because NAT particles move in the vertical, it is also important to characterize the vertical structure of the vortex and cold pool. Figure 2 shows a vertical profile of the separation between the cold pool and vortex centroids on each potential temperature surface for the 0, 5, 10, 15, and 20 rotated temperature fields. There is little vertical variation of centroid separation between the 500 K and 600 K potential temperature surfaces for all rotations of the temperature field. Consequently, it is valid to consider the centroid separation at one particular q level in that range as representative of the vortex as a whole. The profile for the nearly concentric 0 case has a different shape than the other fields because the separation distance is in the meridional direction, while for the others, the distance is in a more zonal direction HNO 3 and H 2 O Fields [22] The initial gas phase nitric acid field was assumed uniform throughout the model domain at 8 ppbv. This value is representative of aircraft measurements made in the lower stratosphere in January 2000 [Kleinböehl et al., 2002]. The initial H 2 O field was also assumed uniform at 5 ppmv. This is also a typical value [see Schiller et al., 2002] Equilibrium and Nonequilibrium Simulations [23] In order to understand better the influence of timedependent particle growth on denitrification, we compared the results of our full nonequilibrium model with a 3-D model in which the NAT particles were assumed to be in thermodynamic equilibrium. [24] The equilibrium model is SLIMCAT, as used as the basis for the nonequilibrium model, but with its own much simpler treatment of NAT particles. Denitrification in the equilibrium model is controlled by fixing a particle size (and hence sedimentation speed) in all grid boxes with temperatures below T NAT and calculating a particle number concentration such that the amount of condensed HNO 3 results in an HNO 3 partial pressure in equilibrium with NAT [Davies et al., 2002]. A NAT particle radius of 2 mm was chosen so that the vortex average denitrification for the equilibrium model 0 run was approximately equal to that of the nonequilibrium model 0 run after 10 days. This particle size is much smaller than the mean size in the nonequilibrium model. Using comparable particle sizes in the equilibrium model would lead to almost complete denitrification at some levels, which is undesirable, since denitrification would then be limited by the total available HNO 3 and therefore make the results difficult to compare with the nonequilibrium model. In these equilibrium model runs, the wind and temperature fields were again held fixed for the duration of each of the 10-day runs. 4. Model Results 4.1. Denitrification Fields [25] Figure 3 shows maps of the volume mixing ratio (VMR) of HNO 3 at q = 545 K for the nonequilibrium model runs with temperature field rotations of 0,10, 15, and 20. The 545 K potential temperature level was chosen because it was the level with maximum vortex-averaged denitrification. Figure 4 shows equivalent maps for the equilibrium model run. In this case, the results refer to the 585 K surface, which has the peak denitrification. The vertical profile of the vortex-averaged denitrification and renitrification is shown as a function of time in Figure 5 and as a function of rotation of the temperature field in Figure 6. The vortex average refers to all air within MPV = 40 PV units. [26] The denitrification at the 545 K level for the nonequilibrium 0 case reaches a maximum of about 3 ppbv near the centroid of the vortex and decreases to zero at the vortex edge (Figure 3). With the cold pool shifted by 10, the peak denitrification is reduced to 2 ppbv, and with a 15 rotation the region with HNO 3 > 1 ppbv has diminished to a small patch in the center of the vortex. With a 20 rotation of the temperature field, approximately half of the NAT region remains within the vortex, but the peak denitrification has reduced to just 0.25 ppbv. These large differences in denitrification occur in response to a change in the configuration of the wind and temperature fields, since the temperature field is identical in each model run. [27] In the 0 equilibrium model run (Figure 4) the peak denitrification of between 3 and 4 ppbv is close to the vortex center, as in the nonequilibrium model run (however, no significance should be attached to the magnitude of

6 AAC 13-6 MANN ET AL.: VORTEX CONCENTRICITY AND ARCTIC DENITRIFICATION Figure 3. HNO 3 VMR on the 545 K potential temperature surface on day 10 of the nonequilibrium model simulations. The different rotations of the temperature field are indicated in each plot (see Figure 1). denitrification, since particle sizes were fixed in order to produce denitrification comparable to the nonequilibrium run). As the temperature field is shifted, the change in predicted denitrification is very different from that in the nonequilibrium model. For the equilibrium model, the area of significant denitrification increases as the vortex-cold pool separation increases, showing that denitrification is occurring equally efficiently in all air below T NAT, irrespective of the configuration of the temperature and wind flow fields. The peak denitrification does decrease as the cold pool is shifted, but this is because the denitrified air is distributed over a larger area. [28] Figure 5 shows the temporal variation of the vertical profiles of vortex-averaged denitrification for the nonequilibrium and equilibrium 0 simulations. In the nonequilibrium simulation (Figure 5a), denitrification begins initially around 590 K after day 2 and then grows to a pronounced region of significant denitrification by day 10 between 450 and 600 K with a maximum vortex-averaged denitrification at 545 K. By day 10 there is a significant region of renitrification between 400 and 450 K where the large denitrifying NAT particles evaporate in warmer air. [29] It is interesting to note that the magnitude of denitrification decreases with decreasing altitude, even though the HNO 3 mixing ratio is constant with altitude. Popp et al. [2001] observed that the magnitude of denitrification matched the profile of available HNO 3 in their observations in January Our results show that such a straightforward interpretation of the denitrification profile will not always be appropriate. The change in denitrification with decreasing altitude below 545 K in our model is caused by a balance between increasing particle size (leading to more efficient denitrification) and an increasing particle loss rate by evaporation (causing renitrification). This balance is discussed further in section 4.2. [30] The vertical profile of HNO 3 in the equilibrium model (Figure 5b) is quite different from the nonequilibrium case. First, the denitrification begins much higher up in the equilibrium model, at around q = 650 K. This is due to the fact that in the equilibrium model, the particles begin sedimenting at a velocity governed by the Stokes drag on a2-mm particle as soon as it is nucleated. No account is taken of the time required to grow to that size, since the assumption in this model is that all particle phase HNO 3 is in the form of NAT particles of radius 2 mm. Second, the shape of the denitrification region is different. The denitrified section of the vortex-averaged profiles from the equilibrium model runs is symmetric about the maximum denitrification, while the nonequilibrium model runs have an asymmetric profile. This is caused by there being no increase in particle size in the equilibrium model; so the first of the two competing effects described above is not present. For the equilibrium model runs, the shape is soleley governed by where and over how wide an area the temper-

7 MANN ET AL.: VORTEX CONCENTRICITY AND ARCTIC DENITRIFICATION AAC 13-7 Figure 4. HNO 3 VMR on the 585 K potential temperature surface on day 10 of the equilibrium model simulations. The different rotations of the temperature field are indicated in each plot (see Figure 1). ature is below T NAT. Third, denitrification remains insignificant for the first 2 days of the nonequilibrium model, while the denitrification starts immediately where T < T NAT in the equilibrium model. This again reflects the time taken for the particles to grow large enough to acquire an appreciable sedimentation speed in the nonequilibrium model. [31] The change in the vertical profile of denitrification with the shift in the cold pool is also interesting to compare in the nonequilibrium and equilibrium runs (Figure 6). The vortex-averaged denitrification in the nonequilibrium run decreases with successively larger temperature field rotations until, at 20 rotation, denitrification is negligible. In Figure 5. Time dependence of the vertical profile of vortex-averaged denitrification in the (a) nonequilibrium and (b) equilibrium simulations. The vortex edge is defined as MPV = 40 PV units.

8 AAC 13-8 MANN ET AL.: VORTEX CONCENTRICITY AND ARCTIC DENITRIFICATION Figure 6. Vertical profiles of vortex-averaged denitrification in the (a) nonequilibrium and (b) equilibrium simulations as a function of the rotation of the temperature field. The vortex edge is defined as MPV = 40 PV units. contrast, in the equilibrium model, the denitrification profile is only marginally reduced for an increased baroclinicity Factors Controlling Denitrification [32] We now describe the way in which the meteorology of the Arctic vortex controls the evolution of the particles and the impact that this has on the calculated denitrification shown above. [33] The Lagrangian nature of the nonequilibrium model enables the behavior of individual particles to be tracked over a complete trajectory from nucleation to complete evaporation. Diagnostics such as the particle lifetime and change in particle altitude with time can be used to understand better the link between meteorology, particle size, and resulting denitrification Particle Lifetimes [34] Figure 7 shows the time between particle nucleation and complete evaporation and how this time varies as the vortex concentricity is changed. The particle lifetimes refer to those particles nucleated during the first day of the simulation and are presented as averages over all such model particles. The particle lifetime is a measure of how well the vortex wind field can keep particles within the cold pool. For the near-concentric 0 case, a bimodal distribution exists with one peak at 6 days and the overall maximum being in the smallest lifetime bin. For this case a significant proportion of particles grow for continuous periods of 8, 9, and even 10 days (these particles still have not completely sublimated by the end of the 10-day simulation). [35] For the model run with a 5 rotated temperature field, the second peak in the bimodal distribution has shifted to around 4.5 days, and the proportion of particles which last longer than 5 days is reduced significantly. For the 10 run the second peak has effectively vanished with the average particle lifetime reduced significantly. For the 20 run, around 99% of the particles have a lifetime of under 3 days, illustrating the controlling influence it has on the denitrification. [36] Note that the particle lifetimes in Figure 7 are shorter than the typical time spent by isentropic air parcels below T NAT [Tabazadeh et al., 2001]. This difference exists because particle growth is often terminated by falling into warm air, thus highlighting the importance of examining particle trajectories rather than air parcel trajectories as a measure of potential denitrification Particle Sizes [37] Related to the particle lifetime is the particle size, which is shown in Figure 8. The size distributions are presented as an average over all model particles existing at the end of day 10. For the 0 and 5 cases, which are both nearconcentric and have a large proportion of particles which exist for a long time, 50% of particles have a radius greater than 4 mm, whereas for the 20 case only 2.5% are in this large size range. In the 20 case, the vast majority of particles are between 0 and 2 mm, which is in accord with their short lifetimes. Figure 7. Distribution of particle lifetimes in the model for different rotations of the temperature field. Lifetimes indicate the time between particle formation and complete evaporation and refer to particles nucleated on the first day of the simulation.

9 MANN ET AL.: VORTEX CONCENTRICITY AND ARCTIC DENITRIFICATION AAC 13-9 Figure 8. Distribution of particle sizes in the model for different rotations of the temperature field. The average was calculated over all particles in the model at the end of day 10; so it includes all q levels Particle Sedimentation Distances [38] The lifetime of an individual particle determines the size to which it can grow and hence the vertical distance over which it can sediment before it evaporates. Figure 9 shows the vertical distribution of all particles that were nucleated during the first day of the simulation and the height they reach when they eventually evaporate completely. For the near-concentric 0 run, the majority of particles sediment to the bottom of the vortex between 380 and 450 K, leading to denitrification at upper levels and renitrification below. In contrast, for the strongly baroclinic 20 rotated run, the altitude distribution at complete evaporation has only changed a little from that at nucleation, resulting in a much reduced downward transport of HNO 3. Also note the increasing loss rate of particles with decreasing altitude below about 500 K in the concentric case, which is important in determining the profile of denitrification (see section 4.1) The Importance of Vortex Concentricity [39] Together, Figures 7, 8, and 9 show that a shift toward a less concentric vortex state reduces the lifetime and hence size and net downward transport of NAT particles. This, in turn, reduces significantly the amount of denitrification. Thus the relative position of the cold pool and vortex is a key parameter in determining the extent of denitrification. Denitrification by large NAT particles in these simulations is essentially switched off once approximately half of the NAT region remains within the vortex. It is the slow growth of the NAT particles that is responsible for this sensitivity. Allowing the particles to instantaneously reach equilibrium sizes removes this sensitivity entirely. In these simulations, because the cold pool size was kept constant, the equilibrium Figure 9. Vertical distributions of the model particles for different rotations of the temperature field. Dashed line, distribution of initial particles during the first day of the simulation; solid line, distribution of the same particle altitudes at complete evaporation. Insets show the relative positions of the vortex edge and NAT region.

10 AAC MANN ET AL.: VORTEX CONCENTRICITY AND ARCTIC DENITRIFICATION Figure 10. Vortex-averaged denitrification as a function of the relative location of the cold pool and vortex centers. The relative positions are expressed in terms of the normalized separation of the NAT region and vortex centroids. A normalized separation of 0 indicates that the cold pool and vortex centers are colocated, while a value of 1 indicates that the centroid of the NAT region sits on the vortex edge. The vortex edge is defined as MPV = 40 PV units. model runs produced almost identical total denitrification, regardless of the relative position of the vortex and cold pool. [40] To summarize the sensitivity of denitrification to the relative locations of the vortex and cold pool, Figure 10 shows the vortex-averaged denitrification as a function of the separation of the cold pool and vortex centroids. The vortex-averaged denitrification was calculated at each level for nonequilibrium model runs assuming the vortex edge is given by MPV = 40 PV units. The near-concentric 0 run has the shortest distance between the two centroids and has the maximum denitrification. As the vortex cold pool separation increases, the denitrification decreases dramatically as seen in section 4.1, so that when the vortex cold pool separation is around 1 vortex effective radius for the 20 run, the denitrification is negligible. In contrast, the equilibrium model run shows almost no change in denitrification as the cold pool is shifted. 5. Summary [41] We have presented the first 3-D nonequilibrium model simulations of Arctic denitrification by low number densities of large NAT particles. The model captures the growth and sedimentation of particles which occurs when such particles are not in equilibrium with gas phase HNO 3. Model NAT particle number densities are in the range 10 5 to 10 3 in the lower stratosphere, similar to those observed during the winter 1999/2000 [Fahey et al., 2001] (for direct comparisons, see Carslaw et al. [2002]). [42] Denitrification by low number densities of NAT particles presents a particular modeling challenge because of the long growth times of the particles [Fahey et al., 2001; Carslaw et al., 2002] and, hence, the highly nonlinear dependence of denitrification on the time that individual particles have available for growth [Jensen et al., 2002]. [43] The idealized model simulations shown here used fixed forcing fields over 10 days where the temperature field had been rotated with respect to the vortex to simulate changes in vortex concentricity. These simulations were designed to explore how the meteorology of the Arctic stratosphere controls the growth period of individual particles and how this in turn affects denitrification. We have shown that the growth period of NAT particles is controlled in the Arctic stratosphere by the relative location of the vortex and cold pool. When the cold pool and vortex are colocated, NAT particles can spend long continuous periods below the NAT temperature and therefore reach large sizes. Such concentric, or Antarctic-like, conditions were shown to be optimum for denitrification, since particles are able to grow for continuous periods of 8 days or more. When the cold pool and vortex positions are offset, the period spent below the NAT temperature is shortened. Such nonconcentric or strongly baroclinic states of the Arctic vortex result in a considerable decrease in the amount of denitrification compared with the near-concentric situation. We showed that denitrification could be switched off entirely even with half of the NAT region remaining within the vortex. Under these conditions, NAT particles were able to grow for continuous periods of only 2 days or less and so reached much smaller sizes. [44] The relative location of the cold pool and vortex is a feature of the Arctic stratosphere that displays great variability, both within one winter and interanually (unlike in the Antarctic) [Pawson et al., 1995]. Because of the strong control exerted by the relative location of the cold pool and vortex, it is unlikely that denitrification in different winters will correlate well with the minimum temperature, although the minimum temperature will also be an important parameter, particularly if the frost point is reached. Although a highly concentric vortex state is optimum for denitrification, a less concentric state is optimum for chemical ozone loss [O Neill et al., 1994; Manney et al., 1996]. Further studies are needed to understand the coupling between vortex concentricity, denitrification, and ozone loss during an entire winter/spring season. [45] Ideally, the trajectories of particles need to be considered by using a nonequilibrium growth and sedimentation model in order to predict denitrification. An equilibrium model cannot take account of the varying times available for growth which result from different arrangements of the cold pool in relation to the vortex. An approximate indicator of the amount of denitrification could be provided by the separation of the cold pool and vortex centroids, as shown here. For the case that we considered,

11 MANN ET AL.: VORTEX CONCENTRICITY AND ARCTIC DENITRIFICATION AAC the amount of denitrification decreased approximately linearly as the centroid of the cold pool was shifted from the centroid of the vortex to the edge of the vortex. However, the highly variable shape of the Arctic vortex and cold pool may make this measure of limited use. [46] An analysis of previous winters is needed in order to determine to what extent conditions were optimum for denitrification by large NAT particles. There is also a need to determine how often, and under what conditions, large NAT particles form. [47] Acknowledgments. G.W.M. was funded by the European Union Framework V MAPSCORE project; S.D. was funded by the Natural Environment Research Council. We would also like to thank the British Atmospheric Data Centre for providing the ECMWF analyses. References Carslaw, K. S., B. P. Luo, and T. Peter, An analytic expression for the composition of aqueous HNO 3 -H 2 SO 4 stratospheric aerosols including gas-phase removal of HNO 3, Geophys. Res. Lett., 22, , Carslaw, K. S., J. Kettleborough, M. J. Northway, S. Davies, R.-S. Gao, D. W. Fahey, D. G. Baumgardner, M. P. Chipperfield, and A. Kleinböhl, A vortex-scale simulation of the growth and sedimentation of large nitric acid particles observed during SOLVE/THESEO 2000, J. Geophys. Res., 107(D20), 8300, doi: /2001jd000467, Chipperfield, M. P., Multiannual simulations with a three-dimensional chemical transport model, J. Geophys. Res., 104(D1), , Chipperfield, M. P., and J. A. Pyle, Model sensitivity studies of Arctic ozone depletion, J. Geophs. Res., 103(D21), 28,389 28,403, Chipperfield, M. P., M. L. Santee, L. Froidevaux, G. L. Manney, W. G. Read, J. W. Waters, A. E. Roche, and J. M. Russell, Analysis of UARS data in the southern polar vortex in September 1992 using a chemical transport model, J. Geophys. Res., 101(D13), 18,861 18,881, Considine, D. B., A. R. Douglass, P. S. Connell, D. E. Kinnison, and D. A. Rotman, A polar stratospheric cloud parameterization for the three-dimensional model of the global modeling initiative and its response to stratospheric aircraft emissions, J. Geophys. Res., 105(D3), , Davies, S., et al., Modeling the effect of denitrification on Arctic ozone depletion during winter 1999/2000, J. Geophys. Res., doi: / 2001JD000445, in press, Dessler, A. E., J. Wu, M. L. Santee, and M. R. Schoeberl, Satellite observations of temporary and irreversible denitrification, J. Geophys. Res., 104(D11), 13,993 14,002, Fahey, D. W., D. M. Murphy, K. K. Kelly, M. K. W. Ko, M. H. Proffitt, C. S. Eubank, G. V. Ferry, M. Loewenstein, and K. R. Chan, Measurements of nitric oxide and total reactive nitrogen in the Antarctic stratosphere: Observations and chemical implications, J. Geophys. Res., 94(D14), 16,665 16,681, Fahey, D. W., et al., The detection of large HNO 3 containing particles in the winter Arctic stratosphere and their role in denitrification, Science, 291, , Hanson, D., and K. Mauersberger, Laboratory studies of the nitric acid trihydrate: Implications for the south polar stratosphere, Geophys. Res. Lett., 15, , Hintsa, E. J., et al., Dehydration and denitrification in the Arctic polar vortex during the winter, Geophys. Res. Lett., 25, , Hübler, G., D. W. Fahey, K. K. Kelly, D. D. Montzka, M. A. Carroll, A. F. Tuck, L. E. Heidt, W. H. Pollock, G. L. Gregory, and J. F. Vedder, Redistribution of reactive odd nitrogen in the lower Arctic stratosphere, Geophys. Res. Lett., 17, , Jensen, E., O. B. Toon, K. Drdla, and A. Tabazadeh, Impact of polar stratospheric cloud particle composition, number density, and lifetime on denitrification, J. Geophys. Res., 107(D20), 8284, doi: / 2001JD000440, Kleinböhl, A., et al., Vortexwide denitrification of the Arctic polar stratosphere in winter 1999/2000 determined by remote observations, J. Geophys. Res., 107(D20), 8305, doi: /2001jd001042, Kondo, Y., et al., NO y -N 2 O correlation observed inside the Arctic vortex in February 1997: Dynamical and chemical effects, J. Geophys. Res., 104(D7), , Kondo, Y., H. Irie, M. Koike, and G. E. Bodeker, Denitrification and nitrification in the Arctic stratosphere during winter of , Geophys. Res. Lett., 27, , Lait, L. R., An alternative form for potential vorticity, J. Atmos. Sci., 51(12), , Manney, G. L., M. L. Santee, L. Froidevaux, J. W. Waters, and R. W. Zurek, Polar vortex conditions during the Arctic winter: Meteorology and MLS ozone, Geophys. Res. Lett., 23, , Northway, M. J., et al., An analysis of large HNO 3 -containing particles sampled in the Arctic stratosphere during the winter of , J. Geophys. Res., 107(D20), 8298, doi: /2001jd001079, O Neill, A., et al., Evolution of the stratosphere during northern winter 1991/92 as diagnosed from U.K. Meteorological Office analyses, J. Atmos. Sci., 51(20), , Pawson, S., and B. Naujokat, The cold winters of the middle 990s in the northern lower stratosphere, J. Geophys. Res., 104(D12), 14,209 14,222, Pawson, S., B. Naujokat, and K. Labitzke, On the polar stratospheric cloud formation potential of the northern stratosphere, J. Geophys. Res., 100(D11), 23,215 23,225, Popp, P. J., et al., Severe and extensive denitrification in the Arctic winter stratosphere, Geophys. Res. Lett., 28, , Pruppacher, H. R., and J. D. Klett, Microphysics of Clouds and Precipitation, 954 pp., Kluwer Acad., Norwell, Mass., Rex, M., et al., Prolonged stratospheric ozone loss in the Arctic winter, Nature, 389, , Santee, M. L., G. L. Manney, L. Froidevaux, W. G. Read, and J. W. Waters, Six years of UARS Microwave Limb Sounder HNO 3 observations: Seasonal, interhemispheric, and interannual differences in the lower stratosphere, J. Geophys., Res., 104(D7), , Santee, M. L., G. L. Manney, N. J. Livesey, and J. W. Waters, UARS Microwave Limb Sounder observations of denitrification and ozone loss in the 2000 Arctic late winter, Geophys. Res. Lett., 27, , Schiller, C., et al., Dehydration in the Arctic stratosphere during the THE- SEO 2000/SOLVE campaigns, J. Geophys. Res., 107(D20), 8293, doi: /2001jd000463, Solomon, S., Stratospheric ozone depletion: A review of concepts and history, Rev. Geophys., 37, , Tabazadeh, A., M. L. Santee, M. Y. Danilin, H. C. Pumphrey, P. A. Newman, P. J. Hamill, and J. L. Mergenthaler, Quantififying denitrification and its effect on ozone recovery, Science, 288, , Tabazadeh, A., E. J. Jensen, O. B. Toon, K. Drdla, and M. R. Schoeberl, Role of the stratospheric polar freezing belt in denitrification, Science, 291, , Waibel, A. E., T. Peter, K. S. Carslaw, H. Oelhaf, G. Wetzel, P. J. Crutzen, U. Poschl, A. Tsias, E. Reimer, and H. Fischer, Arctic ozone loss due to denitrification, Science, 283, , K. S. Carslaw, M. P. Chipperfield, S. Davies, and G. W. Mann, Institute for Atmospheric Science, School of the Environment, University of Leeds, Leeds LS2 9JT, UK. (carslaw@env.leeds.ac.uk; martyn@env.leeds.ac.uk; stewart@env.leeds.ac.uk; gmann@env.leeds.ac.uk) J. Kettleborough, British Atmospheric Data Centre, Rutherford Appleton Laboratory, Chilton, Didcot, Oxon OX11 0QX, UK. (J.A.Kettleborough@ rl.ac.uk)

Measurements of large stratospheric particles in the Arctic polar vortex

Measurements of large stratospheric particles in the Arctic polar vortex JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 108, NO. D20, 4652, doi:10.1029/2002jd003278, 2003 Measurements of large stratospheric particles in the Arctic polar vortex Sarah D. Brooks, 1 Darrel Baumgardner,

More information

Understanding the Relation between V PSC and Arctic Ozone Loss

Understanding the Relation between V PSC and Arctic Ozone Loss Understanding the Relation between V PSC and Arctic Ozone Loss Neil Harris European Ozone Research Coordinating Unit Department of Chemistry, University of Cambridge Ralph Lehmann, Markus Rex, Peter von

More information

Edinburgh Research Explorer

Edinburgh Research Explorer Edinburgh Research Explorer Polar processing and development of the 2004 Antarctic ozone hole: First results from MLS on Aura Citation for published version: Santee, ML, Manney, GL, Livesey, NJ, Froidevaux,

More information

Microphysical modeling of the Arctic winter: 3. Impact of homogeneous freezing on polar stratospheric clouds

Microphysical modeling of the Arctic winter: 3. Impact of homogeneous freezing on polar stratospheric clouds JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 109,, doi:10.1029/2003jd004352, 2004 Microphysical modeling of the 1999--2000 Arctic winter: 3. Impact of homogeneous freezing on polar stratospheric clouds K. Drdla

More information

Hydration, dehydration, and the total hydrogen budget of the 1999/2000 winter Arctic stratosphere

Hydration, dehydration, and the total hydrogen budget of the 1999/2000 winter Arctic stratosphere JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 108, NO. D5, 8320, doi:10.1029/2001jd001257, 2003 Hydration, dehydration, and the total hydrogen budget of the 1999/2000 winter Arctic stratosphere R. L. Herman, 1

More information

JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 108, NO. D5, 8317, doi: /2001jd001063, 2003

JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 108, NO. D5, 8317, doi: /2001jd001063, 2003 JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 108, NO. D5, 8317, doi:10.1029/2001jd001063, 2003 Large-scale chemical evolution of the Arctic vortex during the 1999/ 2000 winter: HALOE/POAM III Lagrangian photochemical

More information

JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 115, D21303, doi: /2009jd012947, 2010

JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 115, D21303, doi: /2009jd012947, 2010 JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 115,, doi:10.1029/2009jd012947, 2010 In situ balloon borne measurements of HNO 3 and HCl stratospheric vertical profiles influenced by polar stratospheric cloud formation

More information

Simulation of polar stratospheric clouds in the specified dynamics version of the whole atmosphere community climate model

Simulation of polar stratospheric clouds in the specified dynamics version of the whole atmosphere community climate model JOURNAL OF GEOPHYSICAL RESEARCH: ATMOSPHERES, VOL. 118, 1 1, doi:./jgrd.1, 1 Simulation of polar stratospheric clouds in the specified dynamics version of the whole atmosphere community climate model T.

More information

Is Antarctic climate most sensitive to ozone depletion in the middle or lower stratosphere?

Is Antarctic climate most sensitive to ozone depletion in the middle or lower stratosphere? Click Here for Full Article GEOPHYSICAL RESEARCH LETTERS, VOL. 34, L22812, doi:10.1029/2007gl031238, 2007 Is Antarctic climate most sensitive to ozone depletion in the middle or lower stratosphere? S.

More information

Liquid ternary aerosols of HNO 3 / H 2 SO 4 / H 2 O in the Arctic tropopause region

Liquid ternary aerosols of HNO 3 / H 2 SO 4 / H 2 O in the Arctic tropopause region GEOPHYSICAL RESEARCH LETTERS, VOL. 31, L01105, doi:10.1029/2003gl018678, 2004 Liquid ternary aerosols of HNO 3 / H 2 SO 4 / H 2 O in the Arctic tropopause region H. Irie, 1 Y. Kondo, 2 M. Koike, 3 N. Takegawa,

More information

Stratospheric ozone loss in the 1996/1997 Arctic winter: Evaluation based on multiple trajectory analysis for double-sounded air parcels by ILAS

Stratospheric ozone loss in the 1996/1997 Arctic winter: Evaluation based on multiple trajectory analysis for double-sounded air parcels by ILAS JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 107, NO. D24, 8210, doi:10.1029/2001jd000615, 2002 Stratospheric ozone loss in the 1996/1997 Arctic winter: Evaluation based on multiple trajectory analysis for double-sounded

More information

JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 107, NO. D20, 8301, doi: /2001jd000999, 2002

JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 107, NO. D20, 8301, doi: /2001jd000999, 2002 JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 107, NO. D20, 8301, doi:10.1029/2001jd000999, 2002 Microphysical mesoscale simulations of polar stratospheric cloud formation constrained by in situ measurements of

More information

High resolution simulation of recent Arctic and Antarctic stratospheric chemical ozone loss compared to observations

High resolution simulation of recent Arctic and Antarctic stratospheric chemical ozone loss compared to observations J Atmos Chem (2006) 55:205 226 DOI 10.1007/s10874-006-9028-8 High resolution simulation of recent Arctic and Antarctic stratospheric chemical ozone loss compared to observations Om Prakash Tripathi Sophie

More information

Chemical Change in the Arctic Vortex During AASE II

Chemical Change in the Arctic Vortex During AASE II Chemical Change in the Arctic Vortex During AASE II Wesley A. Traub, Kenneth W. Jucks, David G. Johnson, and Kelly V. Chance Smithsonian Astrophysical Observatory, Cambridge, Massachusetts Abstract We

More information

Arctic ozone loss deduced from POAM III satellite observations and the SLIMCAT chemical transport model

Arctic ozone loss deduced from POAM III satellite observations and the SLIMCAT chemical transport model Atmos. Chem. Phys., 5, 597 609, 2005 SRef-ID: 1680-7324/acp/2005-5-597 European Geosciences Union Atmospheric Chemistry and Physics 2002 2003 Arctic ozone loss deduced from POAM III satellite observations

More information

The contributions of chemistry and transport to low arctic ozone in March 2011 derived from Aura MLS observations

The contributions of chemistry and transport to low arctic ozone in March 2011 derived from Aura MLS observations JOURNAL OF GEOPHYSICAL RESEARCH: ATMOSPHERES, VOL. 118, 1563 1576, doi:10.2/jgrd.181, 2013 The contributions of chemistry and transport to low arctic ozone in March 2011 derived from Aura observations

More information

The exceptional Arctic winter 2005/06

The exceptional Arctic winter 2005/06 SPARC data assimilation workshop, 2-4 October 2006 The exceptional Arctic winter 2005/06 An example to investigate polar processes using different assimilations systems 1, Gloria Manney 2,3, Steven Pawson

More information

Characteristics of Arctic polar stratospheric clouds in the winter of 1996/1997 inferred from ILAS measurements

Characteristics of Arctic polar stratospheric clouds in the winter of 1996/1997 inferred from ILAS measurements JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 17, NO. D24, 8, doi:1.129/1jd9, 2 Characteristics of Arctic polar stratospheric clouds in the winter of 1996/1997 inferred from ILAS measurements N. Saitoh, 1 S. Hayashida,

More information

Time variations of descent in the Antarctic vortex during the early winter of 1997

Time variations of descent in the Antarctic vortex during the early winter of 1997 JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 109,, doi:10.1029/2004jd004650, 2004 Time variations of descent in the Antarctic vortex during the early winter of 1997 Nozomi Kawamoto Earth Observation Research

More information

Three-dimensional model study of the Arctic ozone loss in 2002/2003 and comparison with 1999/2000 and 2003/2004

Three-dimensional model study of the Arctic ozone loss in 2002/2003 and comparison with 1999/2000 and 2003/2004 Atmos. Chem. Phys., 5, 139 152, 2005 SRef-ID: 1680-7324/acp/2005-5-139 European Geosciences Union Atmospheric Chemistry and Physics Three-dimensional model study of the Arctic ozone loss in 2002/2003 and

More information

Simulation of Polar Ozone Depletion: An Update

Simulation of Polar Ozone Depletion: An Update Simulation of Polar Ozone Depletion: An Update Image taken from www.zmescience.com D. Kinnison (NCAR), S. Solomon (MIT), and J. Bandoro (MIT) February 17, 2015 WACCM Working Group Meeting, Boulder Co.

More information

2. Sketch a plot of R vs. z. Comment on the shape. Explain physically why R(z) has a maximum in the atmospheric column.

2. Sketch a plot of R vs. z. Comment on the shape. Explain physically why R(z) has a maximum in the atmospheric column. 190 PROBLEMS 10. 1 Shape of the ozone layer Consider a beam of solar radiation of wavelength λ propagating downward in the vertical direction with an actinic flux I at the top of the atmosphere. Assume

More information

DANISH METEOROLOGICAL INSTITUTE

DANISH METEOROLOGICAL INSTITUTE DANISH METEOROLOGICAL INSTITUTE SCIENTIFIC REPORT 00-06 Polar Stratospheric Clouds Microphysical and optical models By Niels Larsen COPENHAGEN 2000 Polar Stratospheric Clouds Microphysical and optical

More information

TRANSPORT STUDIES IN THE SUMMER STRATOSPHERE 2003 USING MIPAS OBSERVATIONS

TRANSPORT STUDIES IN THE SUMMER STRATOSPHERE 2003 USING MIPAS OBSERVATIONS TRANSPORT STUDIES IN THE SUMMER STRATOSPHERE 2003 USING MIPAS OBSERVATIONS Y.J. Orsolini (2), W.A. Lahoz (1), A.J. Geer (1) (1) Data Assimilation Research Centre, DARC, University of Reading, UK (2) Norwegian

More information

Meridional structure of the downwelling branch of the BDC Susann Tegtmeier

Meridional structure of the downwelling branch of the BDC Susann Tegtmeier Meridional structure of the downwelling branch of the BDC Susann Tegtmeier Helmholtz Centre for Ocean Research Kiel (GEOMAR), Germany SPARC Brewer-Dobson Circulation Workshop, Grindelwald, June 2012 no

More information

Severe Arctic ozone loss in the winter 2004/2005: observations from ACE-FTS

Severe Arctic ozone loss in the winter 2004/2005: observations from ACE-FTS Click Here for Full Article GEOPHYSICAL RESEARCH LETTERS, VOL. 33, L15801, doi:10.1029/2006gl026752, 2006 Severe Arctic ozone loss in the winter 2004/2005: observations from ACE-FTS J. J. Jin, 1 K. Semeniuk,

More information

Evidence that Nitric Acid Increases Relative Humidity in Low-Temperature Cirrus

Evidence that Nitric Acid Increases Relative Humidity in Low-Temperature Cirrus Supporting Online Material for: Evidence that Nitric Acid Increases Relative Humidity in Low-Temperature Cirrus Clouds R. S. Gao, P. J. Popp, D. W. Fahey, T. P. Marcy, R. L. Herman, E. M. Weinstock, D.

More information

Fundamental differences between Arctic and Antarctic ozone depletion

Fundamental differences between Arctic and Antarctic ozone depletion Fundamental differences between Arctic and Antarctic ozone depletion Susan Solomon 1, Jessica Haskins, Diane J. Ivy, and Flora Min Department of Earth, Atmospheric, and Planetary Sciences, Massachusetts

More information

A method for estimating the extent of denitrification of arctic polar vortex air from tracer-tracer scatter plots

A method for estimating the extent of denitrification of arctic polar vortex air from tracer-tracer scatter plots JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 107, NO. D13, 4169, 10.1029/2001JD001071, 2002 A method for estimating the extent of denitrification of arctic polar vortex air from tracer-tracer scatter plots J.

More information

JOURNAL OF GEOPHYSICAL RESEARCH, VOL.???, XXXX, DOI: /,

JOURNAL OF GEOPHYSICAL RESEARCH, VOL.???, XXXX, DOI: /, JOURNAL OF GEOPHYSICAL RESEARCH, VOL.???, XXXX, DOI:10.1029/, Analysis of ozone loss in the Arctic stratosphere during the late winter and spring of 1997, using the Chemical Species Mapping on Trajectories

More information

Extremely cold and persistent stratospheric Arctic vortex in the winter of

Extremely cold and persistent stratospheric Arctic vortex in the winter of Article Atmospheric Science September 2013 Vol.58 No.25: 3155 3160 doi: 10.1007/s11434-013-5945-5 Extremely cold and persistent stratospheric Arctic vortex in the winter of 2010 2011 HU YongYun 1* & XIA

More information

Mean age of air and transport in a CTM: Comparison of different ECMWF analyses

Mean age of air and transport in a CTM: Comparison of different ECMWF analyses GEOPHYSICAL RESEARCH LETTERS, VOL. 34, L04801, doi:10.1029/2006gl028515, 2007 Mean age of air and transport in a CTM: Comparison of different ECMWF analyses B. M. Monge-Sanz, 1 M. P. Chipperfield, 1 A.

More information

Dynamical Changes in the Arctic and Antarctic Stratosphere During Spring

Dynamical Changes in the Arctic and Antarctic Stratosphere During Spring Dynamical Changes in the Arctic and Antarctic Stratosphere During Spring U. Langematz and M. Kunze Abstract Short- and long-term changes in the intensity and persistence of the Arctic and Antarctic stratospheric

More information

Severe ozone depletion in the cold Arctic winter

Severe ozone depletion in the cold Arctic winter GEOPHYSICAL RESEARCH LETTERS, VOL. 33, L17815, doi:10.1029/2006gl026945, 2006 Severe ozone depletion in the cold Arctic winter 2004 05 M. von Hobe, 1 A. Ulanovsky, 2 C. M. Volk, 3 J.-U. Grooß, 1 S. Tilmes,

More information

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

High initial time sensitivity of medium range forecasting observed for a stratospheric sudden warming GEOPHYSICAL RESEARCH LETTERS, VOL. 37,, doi:10.1029/2010gl044119, 2010 High initial time sensitivity of medium range forecasting observed for a stratospheric sudden warming Yuhji Kuroda 1 Received 27 May

More information

Three-dimensional tracer initialization and general diagnostics using equivalent PV latitude-potential-temperature coordinates

Three-dimensional tracer initialization and general diagnostics using equivalent PV latitude-potential-temperature coordinates Q. J. R. Meteorol. Soc. (1995), 121, pp. 187-210 551.556.44:551.510.534:551.5O9.313.2 Three-dimensional tracer initialization and general diagnostics using equivalent PV latitude-potential-temperature

More information

ATM 507 Lecture 9 Text reading Section 5.7 Problem Set # 2 due Sept. 30 Next Class Tuesday, Sept. 30 Today s topics Polar Stratospheric Chemistry and the Ozone Hole, Required reading: 20 Questions and

More information

Supporting Online Material for

Supporting Online Material for www.sciencemag.org/cgi/content/full/1153966/dc1 Supporting Online Material for The Sensitivity of Polar Ozone Depletion to Proposed Geoengineering Schemes Simone Tilmes,* Rolf Müller, Ross Salawitch *To

More information

Denitrification inside the stratospheric vortex in the winter of by sedimentation of large nitric acid trihydrate particles

Denitrification inside the stratospheric vortex in the winter of by sedimentation of large nitric acid trihydrate particles JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 107, NO. D16, 10.1029/2001JD001015, 2002 Denitrification inside the stratospheric vortex in the winter of 1999 2000 by sedimentation of large nitric acid trihydrate

More information

Polar stratospheric clouds during SOLVE/THESEO: Comparison of lidar observations with in situ measurements

Polar stratospheric clouds during SOLVE/THESEO: Comparison of lidar observations with in situ measurements JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 109,, doi:10.1029/2003jd003463, 2004 Polar stratospheric clouds during SOLVE/THESEO: Comparison of lidar observations with in situ measurements Sarah D. Brooks, 1,2

More information

Model studies on the sensitivity of upper tropospheric chemistry to heterogeneous uptake of HNO 3 on cirrus ice particles

Model studies on the sensitivity of upper tropospheric chemistry to heterogeneous uptake of HNO 3 on cirrus ice particles JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 107, NO. D23, 4696, doi:10.1029/2001jd000735, 2002 Model studies on the sensitivity of upper tropospheric chemistry to heterogeneous uptake of HNO 3 on cirrus ice

More information

JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 117, D05311, doi: /2011jd016789, 2012

JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 117, D05311, doi: /2011jd016789, 2012 JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 117,, doi:10.1029/2011jd016789, 2012 Ozone loss rates in the Arctic winter stratosphere during 1994 2000 derived from POAM II/III and ILAS observations: Implications

More information

Stratospheric ozone loss in the Arctic winters between 2005 and 2013 derived with ACE-FTS measurements

Stratospheric ozone loss in the Arctic winters between 2005 and 2013 derived with ACE-FTS measurements Stratospheric ozone loss in the Arctic winters between 0 and 13 derived with ACE-FTS measurements Debora Griffin 1, Kaley A. Walker 1, 2, Ingo Wohltmann 3, Sandip S. Dhomse 4,, Markus Rex 3, Martyn P.

More information

In situ mountain-wave polar stratospheric cloud measurements: Implications for nitric acid trihydrate formation

In situ mountain-wave polar stratospheric cloud measurements: Implications for nitric acid trihydrate formation JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 108, NO. D5, 8331, doi:10.1029/2001jd001185, 2003 In situ mountain-wave polar stratospheric cloud measurements: Implications for nitric acid trihydrate formation Christiane

More information

JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 111, D11S11, doi: /2005jd006384, 2006

JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 111, D11S11, doi: /2005jd006384, 2006 JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 111,, doi:10.1029/2005jd006384, 2006 Monthly averages of nitrous oxide and ozone for the Northern and Southern Hemisphere high latitudes: A 1-year climatology derived

More information

The Evolution of the Ozone Collar in the Antarctic Lower Stratosphere during Early August 1994

The Evolution of the Ozone Collar in the Antarctic Lower Stratosphere during Early August 1994 402 JOURNAL OF THE ATMOSPHERIC SCIENCES The Evolution of the Ozone Collar in the Antarctic Lower Stratosphere during Early August 1994 ANNARITA MARIOTTI ENEA, Ocean Atmosphere Dynamics Group, Rome, Italy

More information

Effect of zonal asymmetries in stratospheric ozone on simulated Southern Hemisphere climate trends

Effect of zonal asymmetries in stratospheric ozone on simulated Southern Hemisphere climate trends Click Here for Full Article GEOPHYSICAL RESEARCH LETTERS, VOL. 36, L18701, doi:10.1029/2009gl040419, 2009 Effect of zonal asymmetries in stratospheric ozone on simulated Southern Hemisphere climate trends

More information

10. Stratospheric chemistry. Daniel J. Jacob, Atmospheric Chemistry, Harvard University, Spring 2017

10. Stratospheric chemistry. Daniel J. Jacob, Atmospheric Chemistry, Harvard University, Spring 2017 10. Stratospheric chemistry Daniel J. Jacob, Atmospheric Chemistry, Harvard University, Spring 2017 The ozone layer Dobson unit: physical thickness (0.01 mm) of ozone layer if compressed to 1 atm, 0 o

More information

Aura Microwave Limb Sounder observations of dynamics and transport during the record-breaking 2009 Arctic stratospheric major warming

Aura Microwave Limb Sounder observations of dynamics and transport during the record-breaking 2009 Arctic stratospheric major warming Click Here for Full Article GEOPHYSICAL RESEARCH LETTERS, VOL. 36, L12815, doi:10.1029/2009gl038586, 2009 Aura Microwave Limb Sounder observations of dynamics and transport during the record-breaking 2009

More information

Polar vortex dynamics observed by means of stratospheric and mesospheric CO ground-based measurements carried out at Thule (76.5 N, 68.

Polar vortex dynamics observed by means of stratospheric and mesospheric CO ground-based measurements carried out at Thule (76.5 N, 68. Polar vortex dynamics observed by means of stratospheric and mesospheric CO ground-based measurements carried out at Thule (76.5 N, 68.8 W), Greenland I.Fiorucci 1, G. Muscari 1, P. P. Bertagnolio 1, C.

More information

Atmospheric Chemistry and Physics

Atmospheric Chemistry and Physics Atmos. Chem. Phys., 5, 739 753, 5 www.atmos-chem-phys.org/acp/5/739/ SRef-ID: 8-73/acp/5-5-739 European Geosciences Union Atmospheric Chemistry and Physics Influence of mountain waves and NAT nucleation

More information

On the accuracy of analysed low temperatures in the stratosphere

On the accuracy of analysed low temperatures in the stratosphere Atmospheric Chemistry and Physics On the accuracy of analysed low temperatures in the stratosphere B. M. Knudsen Danish Meteorological Institute, Lyngbyvej 100, 2100 Copenhagen, Denmark Received: 20 June

More information

HIRDLS observations and simulation of a lower stratospheric intrusion of tropical air to high latitudes

HIRDLS observations and simulation of a lower stratospheric intrusion of tropical air to high latitudes Click Here for Full Article GEOPHYSICAL RESEARCH LETTERS, VOL. 35, L21813, doi:10.1029/2008gl035514, 2008 HIRDLS observations and simulation of a lower stratospheric intrusion of tropical air to high latitudes

More information

Simulation of ozone depletion in spring 2000 with the Chemical Lagrangian Model of the Stratosphere (CLaMS)

Simulation of ozone depletion in spring 2000 with the Chemical Lagrangian Model of the Stratosphere (CLaMS) JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 107, NO. D20, 8295, doi:10.1029/2001jd000456, 2002 Simulation of ozone depletion in spring 2000 with the Chemical Lagrangian Model of the Stratosphere (CLaMS) J.-U.

More information

Vertical profiles of activated ClO and ozone loss in the Arctic vortex in January and March 2000: In situ observations and model simulations

Vertical profiles of activated ClO and ozone loss in the Arctic vortex in January and March 2000: In situ observations and model simulations JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 108, NO. D22, 8334, doi:10.1029/2002jd002564, 2003 Vertical profiles of activated ClO and ozone loss in the Arctic vortex in January and March 2000: In situ observations

More information

An Examination of Anomalously Low Column Ozone in the Southern Hemisphere Midlatitudes During 1997

An Examination of Anomalously Low Column Ozone in the Southern Hemisphere Midlatitudes During 1997 San Jose State University From the SelectedWorks of Eugene C. Cordero April, 2002 An Examination of Anomalously Low Column Ozone in the Southern Hemisphere Midlatitudes During 1997 Eugene C. Cordero, San

More information

Traveling planetary-scale Rossby waves in the winter stratosphere: The role of tropospheric baroclinic instability

Traveling planetary-scale Rossby waves in the winter stratosphere: The role of tropospheric baroclinic instability GEOPHYSICAL RESEARCH LETTERS, VOL. 39,, doi:10.1029/2012gl053684, 2012 Traveling planetary-scale Rossby waves in the winter stratosphere: The role of tropospheric baroclinic instability Daniela I. V. Domeisen

More information

Continuous Lidar Monitoring of Polar Stratospheric Clouds at the South Pole

Continuous Lidar Monitoring of Polar Stratospheric Clouds at the South Pole IN BOX INSIGHTS and INNOVATIONS Continuous Lidar Monitoring of Polar Stratospheric Clouds at the South Pole BY J A M E S R. C A M P B E L L, E L L S W O R T H J. W E L T O N, A N D J A M E S D. P olar

More information

PROBLEMS Sources of CO Sources of tropospheric ozone

PROBLEMS Sources of CO Sources of tropospheric ozone 220 PROBLEMS 11. 1 Sources of CO The two principal sources of CO to the atmosphere are oxidation of CH 4 and combustion. Mean rate constants for oxidation of CH 4 and CO by OH in the troposphere are k

More information

Climatology of polar stratospheric clouds based on lidar observations from 1993 to 2001 over McMurdo Station, Antarctica

Climatology of polar stratospheric clouds based on lidar observations from 1993 to 2001 over McMurdo Station, Antarctica JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 109,, doi:10.1029/2004jd004800, 2004 Climatology of polar stratospheric clouds based on lidar observations from 1993 to 2001 over McMurdo Station, Antarctica Alberto

More information

J. Alfred, M. Fromm, R. Bevilacqua, G. Nedoluha, A. Strawa, L. Poole, J. Wickert. To cite this version:

J. Alfred, M. Fromm, R. Bevilacqua, G. Nedoluha, A. Strawa, L. Poole, J. Wickert. To cite this version: Observations and analysis of polar stratospheric clouds detected by POAM III and SAGE III during the SOLVE II/VINTERSOL campaign in the 22/23 Northern Hemisphere winter J. Alfred, M. Fromm, R. Bevilacqua,

More information

Persistent shift of the Arctic polar vortex towards the Eurasian continent in recent decades

Persistent shift of the Arctic polar vortex towards the Eurasian continent in recent decades SUPPLEMENTARY INFORMATION DOI: 10.1038/NCLIMATE3136 Persistent shift of the Arctic polar vortex towards the Eurasian continent in recent decades Jiankai Zhang 1, Wenshou Tian 1 *, Martyn P. Chipperfield

More information

Three-Dimensional Model Study of the Antarctic Ozone Hole in 2002 and Comparison with 2000

Three-Dimensional Model Study of the Antarctic Ozone Hole in 2002 and Comparison with 2000 822 J O U R N A L O F T H E A T M O S P H E R I C S C I E N C E S VOLUME 62 Three-Dimensional Model Study of the Antarctic Ozone Hole in 2002 and Comparison with 2000 W. FENG AND M. P. CHIPPERFIELD Institute

More information

Simulation of Polar Ozone Depletion in SD-WACCM4 / MERRA

Simulation of Polar Ozone Depletion in SD-WACCM4 / MERRA Simulation of Polar Ozone Depletion in SD-WACCM4 / MERRA D. Kinnison (NCAR), S. Solomon (MIT), J. Bandoro (MIT), and R. Garcia (NCAR) June 16, 2015 WACCM Working Group Meeting, Baltimore MD. Image courtesy

More information

Seasonal cycles of O 3, CO, and convective outflow at the tropical tropopause

Seasonal cycles of O 3, CO, and convective outflow at the tropical tropopause GEOPHYSICAL RESEARCH LETTERS, VOL. 33, L16802, doi:10.1029/2006gl026602, 2006 Seasonal cycles of O 3, CO, and convective outflow at the tropical tropopause Ian Folkins, 1 P. Bernath, 2 C. Boone, 2 G. Lesins,

More information

Department of Physics and Astronomy, University of Wyoming, Laramie

Department of Physics and Astronomy, University of Wyoming, Laramie JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 96, NO. D2, PAGES 2897-2912, FEBRUARY 20, 1991 STRATOSPHERIC CLOUD OBSERVATIONS DURING FORMATION OF THE ANTARCTIC OZONE HOLE IN 1989 D. J. Hofmann and T. Deshler Department

More information

Update of the Polar SWIFT model for polar stratospheric ozone loss (Polar SWIFT version 2)

Update of the Polar SWIFT model for polar stratospheric ozone loss (Polar SWIFT version 2) https://doi.org/94/gmd--67-7 Author(s) 7. This work is distributed under the Creative Commons Attribution. License. Update of the Polar SWIFT model for polar stratospheric ozone loss (Polar SWIFT version

More information

Mid High Latitude Cirrus Precipitation Processes. Jon Sauer, Dan Crocker, Yanice Benitez

Mid High Latitude Cirrus Precipitation Processes. Jon Sauer, Dan Crocker, Yanice Benitez Mid High Latitude Cirrus Precipitation Processes Jon Sauer, Dan Crocker, Yanice Benitez Department of Chemistry and Biochemistry, University of California, San Diego, CA 92093, USA *To whom correspondence

More information

On the remarkable Arctic winter in 2008/2009

On the remarkable Arctic winter in 2008/2009 JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 114,, doi:10.1029/2009jd012273, 2009 On the remarkable Arctic winter in 2008/2009 K. Labitzke 1 and M. Kunze 1 Received 17 April 2009; revised 11 June 2009; accepted

More information

Comparison of ClO measurements from the Aura Microwave Limb Sounder to ground-based microwave measurements at Scott Base, Antarctica, in spring 2005

Comparison of ClO measurements from the Aura Microwave Limb Sounder to ground-based microwave measurements at Scott Base, Antarctica, in spring 2005 JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 112,, doi:10.1029/2007jd008792, 2007 Comparison of ClO measurements from the Aura Microwave Limb Sounder to ground-based microwave measurements at Scott Base, Antarctica,

More information

Detailed modeling of mountain wave PSCs

Detailed modeling of mountain wave PSCs Detailed modeling of mountain wave PSCs S. Fueglistaler, S. Buss, B. P. Luo, H. Wernli, H. Flentje, C. A. Hostetler, L. R. Poole, K. S. Carslaw, Th. Peter To cite this version: S. Fueglistaler, S. Buss,

More information

Formation of solid particles in synoptic-scale Arctic PSCs in early winter 2002/2003

Formation of solid particles in synoptic-scale Arctic PSCs in early winter 2002/2003 Formation of solid particles in synoptic-scale Arctic PSCs in early winter 2/3 N. Larsen, B. M. Knudsen, S. H. Svendsen, T. Deshler, J. M. Rosen, R. Kivi, C. Weisser, J. Schreiner, K. Mauerberger, F. Cairo,

More information

Global Warming and Climate Change Part I: Ozone Depletion

Global Warming and Climate Change Part I: Ozone Depletion GCOE-ARS : November 18, 2010 Global Warming and Climate Change Part I: Ozone Depletion YODEN Shigeo Department of Geophysics, Kyoto University 1. Stratospheric Ozone and History of the Earth 2. Observations

More information

Modeling polar ozone loss at the University of Colorado

Modeling polar ozone loss at the University of Colorado Modeling polar ozone loss at the University of Colorado CESM workshop Breckenridge, 06/21/2012 Matthias Brakebusch, Cora E. Randall, Douglas E. Kinnison, Simone Tilmes, Michelle L. Santee, Gloria L. Manney

More information

JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 113, D18305, doi: /2007jd009556, 2008

JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 113, D18305, doi: /2007jd009556, 2008 JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 113,, doi:10.1029/2007jd009556, 2008 Seasonal cycle of averages of nitrous oxide and ozone in the Northern and Southern Hemisphere polar, midlatitude, and tropical

More information

CHAPTER 1. MEASURES OF ATMOSPHERIC COMPOSITION

CHAPTER 1. MEASURES OF ATMOSPHERIC COMPOSITION 1 CHAPTER 1. MEASURES OF ATMOSPHERIC COMPOSITION The objective of atmospheric chemistry is to understand the factors that control the concentrations of chemical species in the atmosphere. In this book

More information

Prediction of cirrus clouds in GCMs

Prediction of cirrus clouds in GCMs Prediction of cirrus clouds in GCMs Bernd Kärcher, Ulrike Burkhardt, Klaus Gierens, and Johannes Hendricks DLR Institut für Physik der Atmosphäre Oberpfaffenhofen, 82234 Wessling, Germany bernd.kaercher@dlr.de

More information

The existence of the edge region of the Antarctic stratospheric vortex

The existence of the edge region of the Antarctic stratospheric vortex JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 117,, doi:10.1029/2011jd015940, 2012 The existence of the edge region of the Antarctic stratospheric vortex Howard K. Roscoe, 1 Wuhu Feng, 2 Martyn P. Chipperfield,

More information

On the Control of the Residual Circulation and Stratospheric Temperatures in the Arctic by Planetary Wave Coupling

On the Control of the Residual Circulation and Stratospheric Temperatures in the Arctic by Planetary Wave Coupling JANUARY 2014 S H A W A N D P E R L W I T Z 195 On the Control of the Residual Circulation and Stratospheric Temperatures in the Arctic by Planetary Wave Coupling TIFFANY A. SHAW Department of Earth and

More information

Aircraft Icing Icing Physics

Aircraft Icing Icing Physics Aircraft Icing Icing Physics Prof. Dr. Dept. Aerospace Engineering, METU Fall 2015 Outline Formation of ice in the atmosphere Supercooled water droplets Mechanism of aircraft icing Icing variations Ice

More information

A three-dimensional model study of long-term mid-high latitude lower stratosphere ozone changes

A three-dimensional model study of long-term mid-high latitude lower stratosphere ozone changes Atmos. Chem. Phys., 3, 1253 1265, 2003 Atmospheric Chemistry and Physics A three-dimensional model study of long-term mid-high latitude lower stratosphere ozone changes M. P. Chipperfield School of the

More information

Chemistry 471/671. Atmospheric Chemistry III: Stratospheric Ozone Depletion

Chemistry 471/671. Atmospheric Chemistry III: Stratospheric Ozone Depletion Chemistry 471/671 Atmospheric Chemistry III: Stratospheric Ozone Depletion 2 The Chapman Mechanism O 2 + hn 2 O( 1 D) O( 1 D) + O 2 + M O 3 + M Exothermic O( 1 D) + O 3 2 O 2 O 3 + hn O( 1 D) + O 2 ( 1

More information

Measurements of Ozone. Why is Ozone Important?

Measurements of Ozone. Why is Ozone Important? Anthropogenic Climate Changes CO 2 CFC CH 4 Human production of freons (CFCs) Ozone Hole Depletion Human production of CO2 and CH4 Global Warming Human change of land use Deforestation (from Earth s Climate:

More information

On recent interannual variability of the Arctic winter mesosphere: Implications for tracer descent

On recent interannual variability of the Arctic winter mesosphere: Implications for tracer descent Click Here for Full Article GEOPHYSICAL RESEARCH LETTERS, VOL. 34, L09806, doi:10.1029/2007gl029293, 2007 On recent interannual variability of the Arctic winter mesosphere: Implications for tracer descent

More information

Direct effects of particle precipitation and ion chemistry in the middle atmosphere

Direct effects of particle precipitation and ion chemistry in the middle atmosphere Direct effects of particle precipitation and ion chemistry in the middle atmosphere P. T. Verronen Finnish Meteorological Institute, Earth Observation Helsinki, Finland Contents of presentation 1. Middle

More information

DANISH METEOROLOGICAL INSTITUTE

DANISH METEOROLOGICAL INSTITUTE DANISH METEOROLOGICAL INSTITUTE SCIENTIFIC REPORT -9 Effects from high-speed civil traffic aircraft emissions on polar stratospheric clouds By Niels Larsen Bjørn M. Knudsen Michael Gauss Giovanni Pitari

More information

Microwave Limb Sounder Observations of Polar Middle Atmosphere: Decadal and Inter-annual Variability

Microwave Limb Sounder Observations of Polar Middle Atmosphere: Decadal and Inter-annual Variability Microwave Limb Sounder Observations of Polar Middle Atmosphere: Decadal and Inter-annual Variability Jae N. Lee 1, Dong L. Wu 2, Alexander ozone Ruzmaikin 1, Gloria J. Manney 1, and Sultan Hameed 4 1.

More information

Meteorological conditions of the stratosphere for the CRISTA 2 campaign, August 1997

Meteorological conditions of the stratosphere for the CRISTA 2 campaign, August 1997 JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 107, NO. D23, 8084, doi:10.1029/2001jd000692, 2002 Meteorological conditions of the stratosphere for the CRISTA 2 campaign, August 1997 G. Günther and D. S. McKenna

More information

WACCM Studies at CU-Boulder

WACCM Studies at CU-Boulder WACCM Studies at CU-Boulder V.L. Harvey, C.E. Randall, O.B. Toon, E. Peck, S. Benze, M. Brakebusch, L. Holt, D. Wheeler, J. France, E. Wolf, Y. Zhu, X. Fang, C. Jackman, M. Mills, D. Marsh Most Topics

More information

Quantifying convective influence on Asian Monsoon UTLS composition using Lagrangian trajectories and Aura MLS observations

Quantifying convective influence on Asian Monsoon UTLS composition using Lagrangian trajectories and Aura MLS observations Quantifying convective influence on Asian Monsoon UTLS composition using Lagrangian trajectories and Aura MLS observations Nathaniel Livesey 1, Leonhard Pfister 2, Michelle Santee 1, William Read 1, Michael

More information

JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 108, NO. D15, 4451, doi: /2002jd002832, 2003

JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 108, NO. D15, 4451, doi: /2002jd002832, 2003 JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 108, NO. D15, 4451, doi:10.1029/2002jd002832, 2003 Law of mass action in the Arctic lower stratospheric polar vortex January March 2000: ClO scaling and the calculation

More information

ACE-FTS measurements across the edge of the winter 2004 Arctic vortex

ACE-FTS measurements across the edge of the winter 2004 Arctic vortex GEOPHYSICAL RESEARCH LETTERS, VOL. 32, L15S05, doi:10.1029/2005gl022671, 2005 ACE-FTS measurements across the edge of the winter 2004 Arctic vortex Ray Nassar, 1 Peter F. Bernath, 1 Chris D. Boone, 1 Gloria

More information

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

Isentropic Analysis. Much of this presentation is due to Jim Moore, SLU Isentropic Analysis Much of this presentation is due to Jim Moore, SLU Utility of Isentropic Analysis Diagnose and visualize vertical motion - through advection of pressure and system-relative flow Depict

More information

The Impact of Polar Stratospheric Ozone Loss on Southern Hemisphere Stratospheric Circulation and Surface Climate

The Impact of Polar Stratospheric Ozone Loss on Southern Hemisphere Stratospheric Circulation and Surface Climate The Impact of Polar Stratospheric Ozone Loss on Southern Hemisphere Stratospheric Circulation and Surface Climate James Keeble, Peter Braesicke, Howard Roscoe and John Pyle James.keeble@atm.ch.cam.ac.uk

More information

Large nitric acid particles at the top of an Arctic stratospheric cloud

Large nitric acid particles at the top of an Arctic stratospheric cloud JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 108, NO. D16, 4517, doi:10.1029/2003jd003479, 2003 Large nitric acid particles at the top of an Arctic stratospheric cloud Terry Deshler, 1 Niels Larsen, 2 Christoph

More information

Temperature. Vertical Thermal Structure. Earth s Climate System. Lecture 1: Introduction to the Climate System

Temperature. Vertical Thermal Structure. Earth s Climate System. Lecture 1: Introduction to the Climate System Lecture 1: Introduction to the Climate System T mass (& radiation) T & mass relation in vertical mass (& energy, weather..) Energy T vertical stability vertical motion thunderstorm What are included in

More information

Downward propagation and statistical forecast of the near-surface weather

Downward propagation and statistical forecast of the near-surface weather JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 110,, doi:10.1029/2004jd005431, 2005 Downward propagation and statistical forecast of the near-surface weather Bo Christiansen Danish Meteorological Institute, Copenhagen,

More information

1 Climatological balances of heat, mass, and angular momentum (and the role of eddies)

1 Climatological balances of heat, mass, and angular momentum (and the role of eddies) 1 Climatological balances of heat, mass, and angular momentum (and the role of eddies) We saw that the middle atmospheric temperature structure (which, through thermal wind balance, determines the mean

More information

Polar stratospheric clouds (PSC) play a primary

Polar stratospheric clouds (PSC) play a primary Continuous Lidar Monitoring of Polar Stratospheric Clouds at the South Pole BY JAMES R. CAMPBELL, ELLSWORTH J. WELTON, AND JAMES D. SPINHIRNE. IN BOX INSIGHTS and INNOVATIONS Polar stratospheric clouds

More information