-
Notifications
You must be signed in to change notification settings - Fork 1
/
spacy_annotations.jsonl
325 lines (325 loc) · 155 KB
/
spacy_annotations.jsonl
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
66
67
68
69
70
71
72
73
74
75
76
77
78
79
80
81
82
83
84
85
86
87
88
89
90
91
92
93
94
95
96
97
98
99
100
101
102
103
104
105
106
107
108
109
110
111
112
113
114
115
116
117
118
119
120
121
122
123
124
125
126
127
128
129
130
131
132
133
134
135
136
137
138
139
140
141
142
143
144
145
146
147
148
149
150
151
152
153
154
155
156
157
158
159
160
161
162
163
164
165
166
167
168
169
170
171
172
173
174
175
176
177
178
179
180
181
182
183
184
185
186
187
188
189
190
191
192
193
194
195
196
197
198
199
200
201
202
203
204
205
206
207
208
209
210
211
212
213
214
215
216
217
218
219
220
221
222
223
224
225
226
227
228
229
230
231
232
233
234
235
236
237
238
239
240
241
242
243
244
245
246
247
248
249
250
251
252
253
254
255
256
257
258
259
260
261
262
263
264
265
266
267
268
269
270
271
272
273
274
275
276
277
278
279
280
281
282
283
284
285
286
287
288
289
290
291
292
293
294
295
296
297
298
299
300
301
302
303
304
305
306
307
308
309
310
311
312
313
314
315
316
317
318
319
320
321
322
323
324
325
["[33] Our main result is that the combination of optimal dynamics, i.e., downward transport following a mid-January SSW, and in situ production of NOx due to several moderate SPEs is not necessarily enough to produce a long-term and clearly NOx-dominated effect on stratospheric ozone. Particularly in the case of 2012, (1) the SPEs were only medium size in strength and did not produce enough NOx and (2) the overall production of NOx by particle precipitation in the MLT region was low, leading to relatively low concentrations of NOx being transported down to the stratosphere after the SSW. It should be noted that the cases presented in this paper are from an elongated period of relatively low solar activity. In order to fully understand the NOx connection between the MLT and stratosphere and its inuence on stratospheric ozone levels, periods of high solar activity (for high overall energetic particle precipitation) should be investigated as well. For this, the continuation of middle atmosphere measurements of NOx and ozone is essential.",{"entities":[]}]
["Waters, J. W., L. Froidevaux, W. G. Read, G. L. Manhey, L. S. Elson, D. A. Flower, R. F. Jarnot, and R. S. Har wood, Stratospheric CIO and ozone from the Microwave Limb Sounder on the Upper Atmosphere Research Satel lite, Nature, 36, 597-602, 1993. ",{"entities":[[169,192,"INSTRUMENT"],[200,236,"SPACECRAFT"]]}]
["0100200300102030405060ACEFTS smACEFTSASURVMR [ppbv]altitude [km](a)402002040(b)ACEVAL [ppbv]100502002050100(c)(ACEVAL)/mean [%]050100150200s [%](d)K. Strong et al.: Validation of ACE-FTS N2O",{"entities":[[179,186,"INSTRUMENT"]]}]
["in the assimilation performance in the UTLS is the quality of data. As shown in Figure 10, the MLS assimilation successfully reduced biases until they were smaller than observational errors for all models in the UTLS. Comparisons against independent ozonesonde observations also indicated that the MLS assimilation performed well (Figure 9). Despite the different magnitudes of error growth, assimilation results (biasf) were almost the same among models (Figure 10). This suggests that the MLS assimilation successfully reduced the model dependency on ozone analysis not only in the stratosphere but also in the UTLS.",{"entities":[[95,98,"INSTRUMENT"],[298,301,"INSTRUMENT"],[491,494,"INSTRUMENT"]]}]
["The 21-year SAGE-II record is a natural candidate to form the basis of a merged data set using multiple instrument types, and it is used in ve of the seven records described here (Tables 17 and Fig. 1). The conceptually simplest extension is to add a single data set to the SAGE-II record to cover the years following 2005 after which SAGE-II was turned off. The period from 2005 to the present has had many operational satellite instruments (Hassler et al., 2014; Tegtmeier et al., 2013). To date, three single-instrument extensions have been made to the SAGE-II record. These include one case extended with limb-scattered measurements from OSIRIS (Optical Spectrograph and Infrared Imager System) onboard the Odin satellite (2001present) (Bourassa et al., 2014; Adams et al., 2014; Sioris et al., 2014) and two cases using the stellar occultation measurements from GOMOS (Global Ozone Monitoring by Occultation of Stars) onboard the ENVISAT satellite (20022012) (Penckwitt et al., 2015; Kyrl et al.,",{"entities":[[642,648,"INSTRUMENT"],[650,697,"INSTRUMENT"],[711,715,"SPACECRAFT"],[935,942,"SPACECRAFT"]]}]
["FIG. 17. Partial column ozone from the PV proxy over the altitude range from (top row) 600 to 1900 K (above about 22 km), (row 2) 350 to 600 K (from about 1322 km), and (row 3) 350 to 1900 K on the dates listed at the top. (row 4) TOMS column ozone. The color scale for row 3 has been shifted by 50 DU relative to that for the TOMS data to account for the missing ozone from the surface to 350 K in the proxy calculations. White areas denote regions of missing data.",{"entities":[[231,235,"INSTRUMENT"],[327,331,"INSTRUMENT"]]}]
["MIPAS reduced spectral resolution UTLS-1 mode measurements of temperature, O3, HNO3, N2O, H2O and relative humidity over ice: retrievals and comparison to MLS",{"entities":[[0,5,"INSTRUMENT"],[155,158,"INSTRUMENT"]]}]
["Slnfc = AlncVSlnfVT AT where Slnf is calculated from the original covariance matrix in the linear domain, Sf, by generalized Gaussian error propagation as",{"entities":[]}]
["0.68 0.85 0.60 0.80 0.67 0.86 0.84 0.90 0.74 0.81 0.83 0.88 0.63 0.85 0.57 0.82 0.29 0.84 0.57 0.55 0.65 0.59 0.59 0.78 0.84 0.94",{"entities":[]}]
["MLS is sensitive to only long vertical wavelength dis turbances. The observable horizontal scales depend on the direction of the ins(cid:127)rur.(cid:127),ent line of sight (LOS) as well as on the numerical (cid:127)echnique used to compute the variances. Waves propagat;ing in horizontal directions perpendicular to the LOS are most visible, while those traveling parallel to it are strongly attenuated as a re suit of the averaging of positive and negative tempera ture perturbations along the LOS. ",{"entities":[[0,3,"INSTRUMENT"]]}]
["Douglass, A., and S. Kawa (1999), Contrast between 1992 and 1997 highlatitude spring Halogen Occultation Experiment observations of lower stratospheric HCl, J. Geophys. Res., 104, 18,739 18,754.",{"entities":[]}]
["In the tropics the amount of variability relative to the zonal mean value decreases with altitude. At the 147 hPa level the tropical rms deviations are constant with latitude and peak at 4-30 in both January and July. The 215 and 316 hPa data also show a peak in the rms deviations at 30N in July. ",{"entities":[]}]
["CPU demands have necessitated the approximations of equivalent spherical particles, and one-dimensional (1-D) radiative transfer in scattering calculations.",{"entities":[]}]
["The above sampling and binning procedure must be set into the context of the constraints of instrument sampling imposed by the yaw cycle maneuvers. Figure 1 illustrates the spatial-temporal coverage for the MLS temperature data covering the complete 1 November 199127 October 1994 data interval; as such, it is slightly different than the yaw cycle coverage for any given year. Our time intervals for binning of data in local time are centered on the 15th of each month, and span 36 days at latitudes equatorward of 34, and 72 days at higher latitudes. This ensures 24 h of local time coverage at most latitudes between 80 (for some months inadequate local time coverage occurred near the yaw boundaries at 34). At latitudes poleward of 34 the high-latitude sampling during many months (i.e., January, April, June, November in the Southern Hemisphere, and May, July, October, December in the Northern Hemisphere) consists of 36 continuous days such that they overlap most of the same monthdays as the sampling equatorward of 34. On the other hand, there are some months (i.e., February, March in the Southern Hemisphere, and August, September in the Northern Hemisphere) where the 36 days of highlatitude data coverage are continuous, but are centered near the beginning or end of an adjacent month (i.e., slipped 18 days from midmonth). There are also some months (i.e., May and December in the Southern Hemisphere, and November and January in the Northern Hemisphere), where the 15th of the month falls in the gap between yaw cycles, and about half the local times binned together originate in the previous and following yaw cycles. It is these latter months (in one hemisphere or the other) that may be most subject to aliasing of the type discussed in the appendix. However, the reader is reminded that the results described here are climatological in the sense that they represent multiyear averages. We have chosen to analyze the data in this fashion, rather than on a yaw cycle by yaw cycle basis, in order to obtain a data product that spans both hemispheres up to 80 without alternating data gaps between the yaw cycles. The multiyear averaging utilized here should ameliorate the shortcomings associated with this chosen methodology.",{"entities":[[207,210,"INSTRUMENT"]]}]
["A test run including only the chemistry of solid PSC particles at all temperatures but neglecting the effects of liquid polar stratospheric clouds is also presented (SOLIDS#1; solid yellow curves). This case does not produce as much ozone loss as the reference or LIQUIDS#3 cases, but produces much more ozone loss than the LIQUIDS#1 or LIQUIDS#2 cases. It is important to note that Figure 2 shows that the depletions due to liquid or solid PSCs are not additive. For example, in Antarctica, about half as much ozone depletion as in the reference run is obtained when only solid particles are considered, but when only liquids are considered over the full range of temperatures, results are very close to those of the reference run. Thus, the reference run is not a sum of the two; rather liquid or solid PSCs can each drive substantial heterogeneous chemistry. Hence, such tests cannot determine specic percentages of ozone loss attributable to different types of PSCs in the real world, even if the simplied microphysics package of the model is assumed to be realistic (see further discussion below) and denitrication is included in all cases.",{"entities":[]}]
["Zhu, X., Summers, M. E., and Strobel, D. F.: Calculation of CO2 15 micron band atmospheric cooling rates by Curtis matrix interpolation of correlated k-coefcients, J. Geophys. Res., 97, 12 787 12 797, 1992.",{"entities":[]}]
["How do our uncertainty estimates compare with previous studies? The WCCOS simulation by Smit and ASOPOS (2014) of a tropical prole estimate a similar range of PO3 values that also maximize around the tropopause (up to ~17%) due to the dominance of the IB/IM term. The main difference is that this term dominates PO3 values throughout the prole, whereas in our study, the ow rate and conversion efciency uncertainty dominate in the stratosphere. Sterling et al. (2017) present PO3 mean proles for April and October at Hilo, Hawaii, and American Samoa (two SHADOZ sites), which we interpret as similar to the seasonal plots shown in Figure 5. Here we observe that their range of PO3 values is similar to ours, that is, maxima at the tropopause (>10%) and minima elsewhere in the prole (~5%). Sterling et al. (2017) show a relative minimum (maximum) in MAM (SON), similar to that shown for Hanoi and Kuala Lumpur (Figures 5e and 5f). In comparison, midlatitude and high-latitude ozonesonde proles from previous studies show that (1) the overall uncertainty is smaller, that is, less than ~10% throughout the prole, (2) the conversion efciency uncertainty term is a signicant contributor to PO3 throughout the prole, and for some sites like Uccle is the dominant uncertainty term, and (3) the impact of the IB/IM term remains an important contributor to PO3 around the tropopause (McMurdo uncertainty analysis; T. Deshler,",{"entities":[]}]
["riod the numbers of Mix-1, Mix-2 and Mix-2 enhanced particles decreased signicantly compared with the previous periods. The presence of widespread ice PSCs disappeared abruptly after the 21 January when after the major warming temperatures increased above Tice. The fourth and last phase occurred from 22 to 28 January 2010 and was dominated by liquid STS clouds. A detailed description and examples of the PSCs observed by CALIPSO during the Arctic winter 2009/2010 can be found in Pitts et al. (2011).",{"entities":[[424,431,"SPACECRAFT"]]}]
["J. Kuttippurath, M. Sinnhuber, B.-M. Sinnhuber, H. Kullmann, K. Kunzi, and J. Notholt (2005), Rapid meridional transport of tropical airmasses to the Arctic during the major stratospheric warming in January 2003, Atmos. Chem. Phys., 5(23), 1291 1299.",{"entities":[]}]
["To globally evaluate the analyses over the whole atmosphere, we also compared them to the total ozone columns measurements derived from OMI products. We used the retrievals obtained with the KNMI DOAS method (Veefkind et al., 2006; data available from http://www.temis.nl. The OMI-DOAS total ozone columns showed a globally averaged agreement better than 2% with ground-based observations (Balis et al., 2007). The data show no signicant dependence on latitude except for the high latitudes of the Southern Hemisphere (SH), where there is a systematic overestimation of the total ozone value by 35%.",{"entities":[[136,139,"INSTRUMENT"],[277,280,"INSTRUMENT"]]}]
["Kalnay, E., et al. (1996), The NCEP/NCAR 40-year reanalysis project, Bull. Am. Meteorol. Soc., 77 (3), 437471,",{"entities":[]}]
["Laborde, M., et al. (2012), Single Particle Soot Photometer intercomparison at the AIDA chamber, Atmos. Meas. Tech., 5, 30773097,",{"entities":[[28,59,"INSTRUMENT"]]}]
["mospheric ozone (ACP/AMT inter-journal SI). It is a result of the Quadrennial Ozone Symposium 2016, Edinburgh, United Kingdom, 49 September 2016.",{"entities":[]}]
["The paper is organized as follows. First the individual instruments and their datasets are presented along with considerations for data usage and a brief description of how cloud inuence is handled (Sect. 2). Considerations for compiling the climatologies are discussed (Sect. 3.1) before they are compared on the basis of their probability density functions (Sect. 3.2) and seasonal averages (Sect. 3.3). Finally, some conclusions from the comparisons are presented and the general agreement of the different climatologies are discussed (Sect. 4).",{"entities":[]}]
["Newman, P. A., M. R. Schoeberl, R. A. Plumb, and J. E. Rosenfield, Mixing rates calculated from potential vorticity, J. Geophys. Res., 93, 5221-5240, 1988. ",{"entities":[]}]
["Livesey, N., Read, W., Froidevaux, L., Lambert, A., Manney, G., Pumphrey, H., Santee, M., Schwartz, M., Wang, S., Coeld, R., Cuddy, D., Fuller, R., Jarnot, R., Jiang, J., Knosp, B., Stek, P., Wagner, P., and Wu, D.: Aura Microwave Limb Sounder (MLS) Version 3.3 Level 2 data quality and description document., Tech. rep., Jet Propulsion Laboratory, California Institute of Technology, available at: http://mls.jpl.nasa.gov/data/v3-3 data quality document.pdf, last access: 26 August 2011.",{"entities":[[216,220,"SPACECRAFT"],[221,243,"INSTRUMENT"],[245,248,"INSTRUMENT"]]}]
["Knudsen, B. M., Pommereau, J. P., Garnier, A., Nunes-Pinharandra. M., Denis, L., Newman, P., Letrenne, G., and Durand, M.: Accuracy of analyzed stratospheric temperatures in the winter arctic vortex from infra red montgolfer long duration balloon ights, Part II: Results, J. Geophys. Res., 107, D20, doi:10.1029/2001JD001329, 2002.",{"entities":[]}]
["months. The vertical coupling process is as follows: An increased (decreased) planetary wave activity from the troposphere drives increased (decreased) temperatures and a reversal or weakening (strengthening) of eastward zonal ow in the polar winter stratosphere. Accordingly, the polar winter mesosphere shows a clear response with a cooling (warming) as well as a reversed or weaker (stronger) northward wind. The cooling in the upper-mesosphere often slightly precedes the beginning of the warming in the stratosphere. These vertical connections agree well with previous studies by others (e.g. Jacobi et al., 1997, 2003; Walterscheid et al., 2000; Hoffmann et al., 2002; Dowdy et al., 2007), but are more completely (multiple years; more parameters; and higher vertical resolution) revealed by the larger observational dataset in this study.",{"entities":[]}]
[" 10oW 0o 10oE 20oE 30oE 36oN 42oN 48oN 54oN 60oN 66oN a)Ope TpO3 10oW 0o 10oE 20oE 30oE 36oN 42oN 48oN 54oN 60oN 66oN b)Ope TpO3 alb 10oW 0o 10oE 20oE 30oE 36oN 42oN 48oN 54oN 60oN 66oN c) Ope TpO3 alb covar 10Ozone difference [%]40302010010203040T. Mielonen et al.: Towards the retrieval of tropospheric ozone with OMI 681 Table 5. Differences in tropospheric (up to 400 hPa 6 km) ozone abundances between IASI and OMI retrievals for 17 July 2007 in Europe. Four different versions of the algorithm are used: Operational is the operational version, TpO3 refers to a version that uses the TpO3 climatology, TpO3_alb refers to a version that uses the TpO3 climatology and linear albedo in the UV2 channel, and TpO3_alb_ covar_10 refers to a version that uses the TpO3 climatology, linear albedo in the UV2 channel and a climatological a priori covariance matrix. Average difference in Dobson units (Ave diff), average relative difference (Ave rel diff) and average standard deviation (Ave SD) are presented. In addition, the difference between the operational OMI retrieval and the other versions (Ave rel diff with ope) in percent are given.",{"entities":[[389,392,"INSTRUMENT"],[481,485,"INSTRUMENT"],[490,493,"INSTRUMENT"],[1133,1136,"INSTRUMENT"]]}]
["In the global average, for both seasons and all [21] latitudes included, at 46.4 hPa MLS is biased low by 8% and high by 30% above the profile peak (21.5 6.8 hPa). The highest correlation coefficients between MLS and SMR values ranging from 0.54 to 0.87 are found between 46.4 and 14.7 hPa highlighting the fair agreement",{"entities":[[85,88,"INSTRUMENT"],[210,213,"INSTRUMENT"],[218,221,"INSTRUMENT"]]}]
["When the polar vortex forms, the majority of chlorine is present in the form of HCl and the remainder is present in the other important reservoir gas ClONO2 (e.g., Solomon, 1999; Santee et al., 2008). Figure 11 shows the partitioning between the various inorganic chlorine species (Cly). The available amount of Cly is about 2.7 to 3.3 ppb at 54 hPa. The increase of Cly over the winter is due to transport from above. Due to stronger descent in the model in the Northern Hemisphere, Cly increases to values that are about 0.3 ppb higher in the Northern Hemisphere in 2005 than in the Southern Hemisphere in spring 2006, although the initial values are similar. Thus, the chlorine potentially available for ozone depletion in the Northern Hemisphere is increased compared to the Southern Hemisphere. In both hemispheres, the initial mixing ratio of HCl is about 2 ppb at 54 hPa (75 % of Cly) and the initial mixing ratio of ClONO2 is about 0.7 ppb (see also Santee et al., 2008). Cly is produced by photolysis and reaction with O(1D) from chlorouorocarbons (CFCs), hydrochlorouorocarbons (HCFCs) and similar species of mainly anthropogenic origin (e.g., WMO, 2011; Montzka, 2012). The relative partitioning within Cly is approximately constant between 70 and 32 hPa (with only a slight increase in Cly with altitude; see Figs. S14, S38, S62 and S86 of the Supplement), so that most of the following discussion applies to the entire altitude range.",{"entities":[]}]
["Waters, J. W., L. Froidevaux, G. L. Manney, W. G. Road, and L. S. Elson, Lower stratospheric Cl0 and 0s in the ",{"entities":[]}]
["More recent studies of Hinssen and Ambaum [2010] show that 50% of the interannual variability of the northern hemispheric stratosphere is determined by variations in the 100 hPa eddy heat ux, a measure of the PWs energy entering the stratosphere [e.g., Coy et al., 1997; Pawson and Naujokat, 1999; Hinssen and Ambaum, 2010]. Newman et al. [2001] found a positive correlation between the eddy heat ux at 100 hPa averaged between 45N and 75N and the polar cap temperature. Polvani and Waugh [2004] showed that weak eddy heat uxes nearly always precede strong polar vortex events, which is consistent with wave-mean ow interaction theory [e.g., Andrews et al., 1987].",{"entities":[]}]
["Smit, H. G. J., & ASOPOS (Panel for the Assessment of Standard Operating Procedures for Ozonesondes) (2014). Quality assurance and quality control for ozonesonde measurements in GAW, World Meteorological Organization, GAW report #201. Retrieved from http://www.wmo. int/pages/prog/arep/gaw/documents/FINAL_GAW_201_Oct_2014.pdf",{"entities":[]}]
["spheric O3 and HNO3 measurements at Thule, Greenland: an intercomparison with Aura MLS observations. Atmos Meas Tech 6:24412453. doi:10.5194/amt-6-2441-2013",{"entities":[[78,82,"SPACECRAFT"],[83,86,"INSTRUMENT"]]}]
["Wang W, Matthes K, Schmidt T, Neef L. 2013. Recent variability of the tropical tropopause inversion layer. Geophysical Research Letters 40: 63086313.",{"entities":[]}]
["the late 1800s and late 1900s, and by a factor of 2 between the 1950s and 1990s (Wilson et al., 2012; Parrish et al., 2012; Marenco et al., 1994; Staehelin et al., 1994; Volz and Kley, 1988). Many locations around the world monitored ozone in the late 1800s and early 1900s using the semi-quantitative Schnbein ozonoscope (Marenco et al., 1994; Bojkov, 1986). These estimates indicate that surface ozone was much lower in those days compared to modern times, but the uncertainty of the measurements is so great that no accurate estimate can be made of the absolute increase in ozone (see the review by Cooper et al., 2014).",{"entities":[[302,321,"INSTRUMENT"]]}]
["Muscari, G., A. G. di Sarra, R. L. de Zafra, F. Lucci, F. Baordo, F. Angelini, and G. Fiocco (2007), Middle atmospheric O3, CO, N2O, HNO3, and temperature profiles during the warm Arctic winter 20012002, J. Geophys. Res., 112, D14304, doi:10.1029/2006JD007849.",{"entities":[]}]
["forcing: coupling to the troposphereocean response. J. Atmos. Sci. 69, 18411864. http://dx.doi.org/10.1175/JAS-D-11-086.1.",{"entities":[]}]
["Even though the ofine MLS data set and the SD-WACCM simulations display similar structures (see Fig. 9), they differ in magnitude. The ofine MLS data suggest that there is more mesospheric HO2 than predicted by the model, particularly at 0.02 hPa, which requires further investigation.",{"entities":[[22,25,"INSTRUMENT"],[141,144,"INSTRUMENT"]]}]
["Xu, J., H.L. Liu, W. Yuan, A. K. Smith, R. G. Roble, C. J. Mertens, J. M. Russell III, and M. G. Mlynczak (2007), Mesopause structure from Thermosphere, Ionosphere, Mesosphere, Energetics, and Dynamics (TIMED)/Sounding of the Atmosphere Using Broadband Emission Radiometry (SABER) observations, J. Geophys. Res., 112, D09102, doi:10.1029/2006JD007711.",{"entities":[]}]
["Considerable uncertainty persists in laboratory measurements of the rate constant for the Reaction (R4). Some studies (e.g., Kovalenko et al., 2007; von Clarmann et al., 2009) pointed out the necessity of increasing it towards the value from Stimpe et al. (1979) that is about a factor-of-two faster than the rate constant (Sander et al., 2006) used in WACCM4. Recently, the new JPL recommendation (Sander et al., 2011) was updated towards a higher value. However this remains an open issue and the possibility of increasing the agreement between observed data and model results by uploading the rate constant should be taken into account.",{"entities":[]}]
["the vertical profiles of temperature, and the CO and O3 gradients during summer for all latitudes between 30N and 70N in 10 latitude bins. The results are shown in Figure 12. Figure 12a shows that the height of the relative maximum in the temperature above the tropopause is increasing with increasing latitude. Figure 12b shows that the absolute maximum in the CO gradient is a climatological feature over a broad range of latitudes. Figure 12c shows that the relative maxima and minima in the O3 gradient at and above the tropopause, respectively, are only apparent for latitude bands >40N. The heights of the minima in the O3 gradient above the tropopause are found to increase with increasing latitude, similarly to the relative maxima in the temperature profiles. [43] For now, we only focus on the maximum in the absolute CO gradient and investigate if it is also found in the SH and for both JJA and DJF. Figure 13 shows that absolute maxima in the CO gradients occur at the thermal tropopause throughout both NH and SH midlatitudes. The relation breaks down in the polar regions during winter, and in the SH subtropical jet region during summer. The maximum in the absolute value of the CO gradient indicates that the thermal tropopause represents a localized minimum in vertical mixing within the tropopause region. This finding is also supported by Figures 2 and 12c, the latter indicating a localized maximum in the O3 gradient at least at latitudes >40N.",{"entities":[]}]
["The occurrence of mesospheric cooling (MC) during major SSWs has been well established in both hemispheres (Walterscheid etal. 2000; Siskind etal. 2005; de Wit et al. 2015). However, such studies during minor SSWs are sparse and limited to few model simulations (Siskind et al. 2010; Chandran et al. 2013). Recently, Eswaraiah etal. (2016) studied the mesospheric dynamics during 2010 SH minor SSW for the first time. Using simultaneous observations of winds by King Sejong Station (KSS, 62.22S, 58.78W) meteor radar (KSS MR) and temperatures by Microwave Limb Sounder (MLS), they reported zonal wind reversal at 8292 km and mesosphere cooling at 7880km.",{"entities":[[579,601,"INSTRUMENT"],[603,606,"INSTRUMENT"]]}]
["Figure 5. Microwave limb sounder (MLS) (version 3) versus halogen occultation experiment (HALOE) (version 17) H20 data. ",{"entities":[[11,33,"INSTRUMENT"],[35,38,"INSTRUMENT"]]}]
["Minschwaner, K., Carver, R. W., Briegleb, B. P., and Roche, A. E.: Infrared radiative forcing and atmospheric lifetimes of trace species based on observations from UARS, J. Geophys. Res., 103, 2324323253, 1998.",{"entities":[[164,168,"SPACECRAFT"]]}]
["Konopka, P., J.-U. Groo, F. Ploeger, and R. Mller (2009), Annual cycle of horizontal in-mixing into the lower tropical stratosphere, J. Geophys. Res., 114, D19111, doi:10.1029/2009JD011955.",{"entities":[]}]
["the mesosphere and stratosphere have been reported as a result of SPEs and the indirect EEP impact on ozone19,21.",{"entities":[]}]
["The shape of the difference prole for the comparison with ACE-MAESTRO is quite similar, but ACE-MAESTRO shows a larger negative bias with respect to the Eureka DIAL observations. Mean relative difference values range from 20 to +7% (on average 13%) in the range 1238 km. The de-biased standard deviation of the mean relative differences is within 10% between 19 and 30 km and increases above",{"entities":[[153,164,"INSTRUMENT"]]}]
["Gille, J. C., et al. (2008), The High Resolution Dynamics Limb Sounder (HIRDLS): Experiment overview, results, and temperature validation, J. Geophys. Res., doi:10.1029/2007JD008824, in press.",{"entities":[[33,70,"INSTRUMENT"],[72,78,"INSTRUMENT"]]}]
["VOmel, H., M. Rummukainen, R. Kivi, J. Karhu, T Turunen, E. Kyr6, J. Rosen, N. Kjome, and S. Oltmans, Dehydration and sedimentation of ice particles in the Arctic stratospheric vortex, Geophys. Res. Lett., 24, 795-798, 1997. ",{"entities":[]}]
["Key Points: (cid:129) Tropical stratospheric upwelling is quantied using water vapor and carbon monoxide measurements",{"entities":[]}]
["compare SCIAMACHY total columns retrieved with the same algorithm to independent data. Eskes et al. [2005] compared TOSOMI data taken during 2003 and 2004 with ground-based measurements at seventy ground locations from Dobson, Brewer, and Filter instruments. They found that SCIAMACHY data have a bias of 1.7%, which is remarkably close to our estimated mean SCIAMACHY innovation.",{"entities":[[8,17,"INSTRUMENT"],[275,284,"INSTRUMENT"],[359,368,"INSTRUMENT"]]}]
["Fig. 6. The retrieved ClO vertical prole in ppb from the TELIS balloon ight on 24 January 2010 in chlorine activated Arctic air. ClO is a member of the active chlorine chemical family and the peak in the prole around 23 km is because of the conversion of inactive reservoir chlorine into active chlorine. The retrieved prole by TELIS is given in black and the dashed lines refer to the estimated overall accuracy of this prole. The prole by the MLS satellite instrument is given in blue. This prole has been weighted by the averaging kernel of the TELIS retrievals.",{"entities":[[57,62,"INSTRUMENT"],[328,333,"INSTRUMENT"],[445,448,"INSTRUMENT"],[548,553,"INSTRUMENT"]]}]
["where respect to the constrained quantities temperature and pressure) and quantities as reported by (6) in the earlier phase.",{"entities":[]}]
["We focus on SH winter months JulyAugustSeptember for 3555 S because the number of proles obtained at midlatitudes during each winter is similar over the life of the mission (see Sect. 2.2). As in the NH, we compare the differences between HALOE and simulated CH4 for 19921998 with 19992005 by examining histograms of the percentage differences for each time period at 46.4, 31.6 and 21.5 hPa",{"entities":[]}]
["[37] The observed discrepancies between the MIPAS and GPS-RO temperatures are also related to the measurement limitation of individual instruments. A detailed analysis of the error budgets has been reported by von Clarmann et al. [2003c] for the MIPAS temperatures and by Hajj et al. [2002] and Wickert et al. [2004a] for the GPS-RO retrievals. They concluded that, for a single profile, the MIPAS temperatures are retrieved with a total error of 0.5 1.5 K at stratospheric altitudes and the GPS-RO temperatures are derived with an accuracy of typically better than 1 K. For a large ensemble, random error components should average out, but the systematic components do not.",{"entities":[[246,251,"INSTRUMENT"],[392,397,"INSTRUMENT"]]}]
["Seasonal cycles at SH high latitudes have a more complex structure than those at the NH mid-latitudes due to generally weaker downwelling in the BrewerDobson circulation and the inuence of Antarctic ozone depletion. As a consequence, the reanalyses have more difculty in capturing the seasonal cycle. At 10 hPa, MERRA-2 shows the best agreement with the observations. CFSR also follows the observations relatively well but overestimates the amplitude of the seasonal cycle, primarily because of values that are too low during May through July. MERRA and JRA-25 are outliers in that they do not contain the strong annual minimum observed during late austral autumn and early winter. At 50 hPa, MERRA and JRA-25 agree better with observations than at 10 hPa, but still underestimate austral springtime ozone depletion. Finally, at 150 hPa, the seasonality in the reanalyses varies widely and is inconsistent with that in the observations, with the exception of MERRA, which produces the most realistic seasonal cycle amplitude. MERRA-2 shows the closest agreement with observations at all levels except for 150 hPa, which is the next to lowest valid level of the MLS v2.2 ozone retrievals that it assimilates.",{"entities":[]}]
["Straub, C., Tschanz, B., Hocke, K., Kmpfer, N., and Smith, A. K.: Transport of mesospheric H2O during and after the stratospheric sudden warming of January 2010: observation and simulation, Atmos. Chem. Phys., 12, 54135427, doi:10.5194/acp-12-54132012, 2012.",{"entities":[]}]
["Bnisch, H., A. Engel, J. Curtius, T. Birner, and P. Hoor (2009), Quantifying transport into the lowermost stratosphere using simultaneous in situ measurements of SF6 and CO2, Atmos. Chem. Phys., 9, 59055919.",{"entities":[]}]
["Figure 3. (a) ERA-Interim zonal mean temperature dierence between 90 and 60 N and time mean removed zonal mean (b) Aura-MLS H2O VMR and (c) MIPAS-ENVISAT CH4 VMR over the equator during 1 December 200818 February 2009.",{"entities":[[115,119,"SPACECRAFT"],[120,123,"INSTRUMENT"],[140,145,"INSTRUMENT"],[146,153,"SPACECRAFT"]]}]
["noise. Comparison with the scatter bars in Figure 2 shows that by far the largest part of the variability seen must be of atmospheric origin.",{"entities":[]}]
["Fig. 4. Time series of (a) daily mean temperature anomalies observed using MLS during November 2008 to March 2009 at 42 km altitude level. Intrinsic mode function components from the rst to fourth IMFs are shown in (b) to (e). Corresponding L-S periodograms are shown in (f) to (j), respectively. Dashed horizontal line in (g) to (j) indicates 95 % condence level.",{"entities":[[75,78,"INSTRUMENT"]]}]
["Akmaev, R.A., Fomichev, V.I., Zhu, X., 2006. Impact of middle-atmospheric composition changes on greenhouse cooling in the upper atmosphere. J. Atmos. Sol. Terr. Phys. 68, 18791889.",{"entities":[]}]
["the SOFIE versus ACE coincidences over full years from 2008 to 2012. The higher coincidence numbers correspond well with what Figure 2a has indicated about the latitude and local time coincidences. One exception is that in the NH in January there are no coincidences given the close latitude and time overlaps because the longitude differences fall out of the 20 interval during this time. Figure 3b shows SOFIE versus MIPAS coincidence numbers. Since MIPAS covers almost exclusively two local times (Figure 2c) while SOFIE local time coverage varies signicantly throughout a year, it is not surprising that the high coincidence numbers occur primarily in the middle part of a year from April to September.",{"entities":[[419,424,"INSTRUMENT"],[452,457,"INSTRUMENT"]]}]
["Russell III, J. M., Mlynczak, M. G., Gordley, L. L., Tansock, J., and Esplin, R.: An overview of the SABER experiment and preliminary calibration results, Proc. SPIE Int. Soc. Opt. Eng., 3756, 277288, 1999.",{"entities":[]}]
["Acknowledgements. This research was supported by NASA grant NNX08AN78G to NMT. Work at the Jet Propulsion Laboratory, California Institute of Technology was done under contract with the National Aeronautics and Space Administration. Herb Picketts role as PI of the MLS OH measurements is acknowledged; the work presented here was made possible only through Herbs substantial accomplishments in instrument design and construction, retrieval algorithms, and data validation. We thank Martyn Chippereld of Leeds University, UK, who developed the SLIMCAT model, for making the code available to us. The useful comments and suggestions of two anonymous referees is acknowledged.",{"entities":[[265,268,"INSTRUMENT"]]}]
["Large variability in Antarctic ozone loss has been seen in the last few years (20042013) relative to other winters since 1992 (e.g., Tilmes et al., 2006; Huck et al., 2007; Yang et al., 2006; Santee et al., 2008a). For instance, the winters of 2004, 2010, and 2012 were relatively warm, with minor warmings and, hence, limited ozone loss (Santee et al., 2005; de Laat and van Weele, 2011; WMO, 2014). The rst fortnight of August 2005 was unusually cold and showed a high rate of ozone loss and an unprecedented ozone hole (WMO, 2014). The winter of 2006 was one of the coldest and, hence, the Antarctic vortex experienced the largest ozone hole to date (Santee et al., 2011; WMO, 2014). The winters of 2007, 2009, and 2013 were characterized by average temperatures and, hence, ozone holes of a moderate size (Tully et al., 2008; Kuttippurath et al., 2013). However, the winters of 2008 and 2011 were again very cold and characterized by large ozone holes (Tully et al., 2011; WMO, 2014). Here, we provide a detailed view of these 10 winters in relation to polar processing and the chemistry of ozone loss. In this study, we discuss (i) the interannual variability in ozone loss and chlorine activation and (ii) horizontal, vertical, and seasonal variability in ozone loss in the Antarctic stratosphere during these (2004 2013) winters. We use high-resolution simulations for analyzing the polar processing and interannual changes in ozone loss in detail. Note that the simulations are highly resolved in the lower stratosphere (about 0.5 km between 425 and 550 K) too, to closely study the ozone loss features in those peak ozone loss altitude layers. Additionally, the past 10 winters offer a good opportunity to test the chemical and dynamical processes in numerical models. Furthermore, observations from the Aura microwave limb sounder (MLS) (Froidevaux et al., 2008; Santee et al., 2008b), one of the best satellite instruments currently available for sampling polar vortices, are compared to the model results. Therefore, for the rst time ozone loss and chlorine activation can be studied with high-resolution measurements that have very good spatial and temporal coverage inside the Antarctic vortex. Previous satellite measurements were relatively limited to a small temporal and spatial area as far as high-latitude observations are concerned (e.g., Tilmes et al., 2006; Hoppel et al., 2005). While the Upper Atmosphere Research Satellite MLS (Waters et al., 1999) had a similar latitudinal coverage, the frequency of its polar measurements was lower than that of Aura MLS (e.g., Livesey et al., 2013; Froidevaux et al., 2008). Therefore, the study with high-resolution simulations (both horizontally and vertically) together with the high-resolution",{"entities":[[1813,1817,"SPACECRAFT"],[1818,1840,"INSTRUMENT"],[1842,1845,"INSTRUMENT"],[2413,2448,"SPACECRAFT"],[2449,2452,"INSTRUMENT"],[2574,2578,"SPACECRAFT"],[2579,2582,"INSTRUMENT"]]}]
["The CLAES July 10, 1992, in Figure 16d provides a classic example of south polar winter HNO3 depletion due to freeze out on PSCs. There is a large maximum in the so-called collar region near 62S and at 30 mbar that is unique to the south polar winter. Similar to the CLAES 93 and the LIMS, it shows a prominent positive horizontal gradient poleward from about 50S at altitudes near 10 mbar and above. In the region from the equator to 32N it more closely resembles the relatively enhanced CLAES January 10, 1992, equator to 32S case than the other two north polar winter examples. ",{"entities":[[4,9,"INSTRUMENT"],[272,277,"INSTRUMENT"],[289,293,"INSTRUMENT"],[501,506,"INSTRUMENT"]]}]
["Schoeberl, M. R., A. R. Douglass, R. S. Stolarski, S. Pawson, S. E. Strahan, and W. Read (2008), Comparison of lower stratospheric",{"entities":[]}]
["and precipitation from radar reectivity and background meteorological data. It reasonably isolates deep convection from trailing high clouds in a complex convective system (Sassen & Wang, 2008); thus, it is useful data for capturing overshooting convection. Only proles classied as deep convection is used for this study. Due to its frequent sampling, horizontal resolution of the cloud sample is ~1 km along track. We dene separate events of deep convection by connecting CloudSat samples located within 100 km along the orbit track. To avoid overlap between deep convection samples (i.e., sampling the same system multiple times), we further thin out the samples to have distance between separate events larger than 1,000 km. Among these data we focus on deep convection groups with maximum cloud top higher than 17 km. The selection method is utilized to avoid duplicated use of the dense along-track observations, and overall results are not sensitive to the selection criteria. One relevant aspect of the CloudSat sampling is that measurements are made near local times of 1:30 a.m. and 1:30 p.m. and hence probably undersample extreme convection over land where a large diurnal cycle occurs, with extreme convection in late afternoon (Liu & Zipser, 2008).",{"entities":[[473,481,"SPACECRAFT"],[1010,1018,"SPACECRAFT"]]}]
["Portmann, R. W., Solomon, S., Fishman, J., Olson, J. R., Kiehl, J. T., and Briegleb, B.: Radiative forcing of the Earths climate system due to tropical tropospheric ozone production, J. Geophys. Res., 102, 94099417, 1997.",{"entities":[]}]
["Oct 1978present continuous coverage; 19701976 NOAA-4 coverage, partial coverage after mid-1972. GOME: 19952003 GOME-2A: 2006present GOME-2B: 2012present 2004present 2007present",{"entities":[]}]
["GWs can propagate both eastward and westward, but only against the zonal ow, implying the presence of eastwardpropagating GWs during summer and westward-propagating GWs during winter. The extratropical mesoand stratospheric zonal winds are very weak and change direction during the equinoxes, resulting in a damping of both westwardand eastward-propagating GWs during these periods (Hoffmann et al., 2010). Enhanced PW activity is observed at the same time (Stray et al., 2014). Temperature enhancements after the spring equinox are related to the nal breakdown of the polar vortex or the last stratospheric warming event (Shepherd et al., 2002). Several studies have observed a springtime tongue of westward ow between 85 and 100 km, occurring approximately from day 95 to 120, reecting the nal warming (e.g. Hoffmann et al., 2010; Manson et al., 2002). The nal warming is characterised by forced planetary Rossby waves that exert a strong westward wave drag from the stratosphere up to 100 km.",{"entities":[]}]
["Figure 13. Same as right panel of Fig. 12, but for individual balloon stations. Left panel: Lauder (SH mid-latitudes), middle panel: Hilo (tropics), right panel: Boulder (NH mid-latitudes).",{"entities":[]}]
["World Meteorological Organization (2010), Scientic Assessment of Ozone Depletion: 2010, WMO Global Ozone Research and Monitoring",{"entities":[]}]
["[2] The atmospheric heating caused by the absorption of solar radiation by ozone and molecular oxygen is important in forcing the circulation and tides in the stratosphere, mesosphere, and lower thermosphere. The heating varies seasonally due to variations in the solar zenith angle and in the EarthSun distance. The heating can also vary due to changes in the amount and vertical structure of middle atmosphere ozone.",{"entities":[]}]
["Schwartz, M. J., et al. (2008), Validation of the Aura Microwave Limb Sounder temperature and geopotential height measurements, J. Geophys. Res., 113, D15S11, doi:10.1029/2007JD008783.",{"entities":[[50,54,"SPACECRAFT"],[55,77,"INSTRUMENT"]]}]
["Zhang, Y., Zhang, X., Wang, K., Zhang, Q., Duan, F.-K., and He, K.-B.: Application of WRF/Chem over East Asia: Part II. Model Improvement and Sensitivity Simulations, Atmospheric Environment, 124, 301320, doi:10.1016/j.atmosenv.2015.07.023, 2016b.",{"entities":[]}]
["Abstract. A three-dimensional (3-D) chemical transport model (CTM), SLIMCAT, has been used to quantify the effect of denitrication on ozone loss for the Arctic winter 2004/2005. The simulated HNO3 is found to be highly sensitive to the polar stratospheric cloud (PSC) scheme used in the model. Here the standard SLIMCAT full chemistry model, which uses a thermodynamic equilibrium PSC scheme, overpredicts the ozone loss for Arctic winter 2004/2005 due to the overestimation of denitrication and stronger chlorine activation than observed. A model run with a coupled detailed microphysical denitrication scheme, DLAPSE (Denitrication by Lagrangian Particle Sedimentation), is less denitried than the standard model run and better reproduces the observed HNO3 as measured by Airborne SUbmillimeter Radiometer (ASUR) and Aura Microwave Limb Sounder (MLS) instruments. Overall, denitrication is responsible for a 30 % enhancement in O3 depletion compared with simulations without denitrication for Arctic winter 2004/2005, which is slightly larger than the inferred impact of denitrication on Arctic ozone loss for previous winters from different CTMs simulations. The overestimated denitrication from standard SLIMCAT simulation causes 510 % more ozone loss at 17 km compared with the simulation using the DLAPSE PSC scheme for Arctic winter 2004/2005. The calculated partial column ozone loss from SLIMCAT using the DLAPSE scheme is about 130 DU by mid-March 2005, which compares well with the inferred column ozone loss from ozonesondes and satellite data (12721 DU).",{"entities":[[774,807,"INSTRUMENT"],[809,813,"INSTRUMENT"],[819,823,"SPACECRAFT"],[824,846,"INSTRUMENT"],[848,851,"INSTRUMENT"]]}]
["Thompson, D. W. J. and Solomon, S.: Interpretation of recent Southern Hemisphere climate change, Science, 296, 895899, doi:10.1126/science.1069270, 2002.",{"entities":[]}]
["[10] We use retrievals of cloud fraction, water vapor mixing ratio, temperature, and geopotential height from the AIRS version 5, level 3 daily gridded product (AIRX3STD) between September 2002 and July 2010. Temperature and humidity profiles are used as input to the FuLiou radiative transfer code [Fu and Liou, 1992] to calculate the radiative cooling rates that are used in determining the clearsky convergence profile, as explained in section 3.2. Geopotential heights are used to convert colocated CloudSat retrievals to a common pressure grid.",{"entities":[[114,118,"INSTRUMENT"],[503,511,"SPACECRAFT"]]}]
["Shepherd, T. G., Plummer, D. A., Scinocca, J. F., Hegglin, M. I., Fioletov, V. E., Reader, M. C., Remsberg, E., von Clarmann, T., Wang, H. J.: Reconciliation of halogen-induced ozone loss with the total-column ozone record, Nat. Geosci., 7, 443449, doi:10.1038/ngeo2155, 2014.",{"entities":[]}]
["National Centers for Environmental Prediction/National Weather Service/NOAA/U.S. Department of Commerce: NCEP FNL operational model global tropospheric analyses continuing from July 1999, Research Data Archive at the National Center for Atmospheric Research, Computational and Information Systems Laboratory, Boulder, Colo., https://doi.org/10.5065/D6M043C6, 2000.",{"entities":[]}]
["Figure 7. The particle volumes from the STS-only model (colored circles) compared with ER-2 aircraft observations (black dots) in January and February, 1989 at 5060 hPa [Dye et al., 1992]. The different colors represent different saturation status of HNO3 over the particles.",{"entities":[]}]
["Ann. Geophys., 31, 967981, 2013 www.ann-geophys.net/31/967/2013/ doi:10.5194/angeo-31-967-2013 Author(s) 2013. CC Attribution 3.0 License.",{"entities":[]}]
["Pulido, M., S. Polavarapu, T. G. Shepherd, and J. Thuburn (2011), Estimation of optimal gravity wave parameters for climate models using a genetic algorithm, Q. J. R. Meteorol. Soc., doi:10.1002/qj.932, in press. Randall, C. E., V. L. Harvey, C. S. Singleton, P. F. Bernath, C. D. Boone, and J. U. Kozyra (2006), Enhanced NOx in 2006 linked to strong upper stratospheric Arctic vortex, Geophys. Res. Lett., 33, L18811, doi:10.1029/2006GL027160.",{"entities":[]}]
["Ozone radiative heating in the upper stratosphere is another important forcing for the diurnal tide, and MLS O3 can be used to accurately monitor the variation of this forcing source. The v5 of O3 is greatly im-",{"entities":[[105,108,"INSTRUMENT"]]}]
["Lipson, J. B., Elrod, M. J., Beiderhase, T. W., Molina, L. T., and Molina, M. J.: Temperature dependence of the rate constant and branching ratio for the OH+ ClO reaction, J. Chem. Soc. Faraday T., 93, 26652673, 1997.",{"entities":[]}]
["The remainder of this paper is organized as follows. Section 2 gives a brief description of the GMI model, 7Be source and cross-tropopause ux, and 7Be and ozone observational data sets used for evaluating the model. Section 3 evaluates model results with UT/LS and surface 7Be data. Section 4 assesses cross-tropopause transport of 7Be in different meteorological elds. Section 5 compares the results with previous modeling studies. Section 6 discusses the implications for the impact of STE on tropospheric ozone, followed by the summary and conclusions in Sect. 7.",{"entities":[]}]
["[55] Our comprehensive and detailed results are especially well suited for further comprehensive comparisons with global 3D chemistry climate models, and we plan efforts in that direction. The results can also be important because the information can be used in conjunction with other data sets that do not provide adequate information in terms of local time but which are important because the data have a time span of decades.",{"entities":[]}]
["Four types of spectrometers, having different spectral resolutions and bandwidths, are used to cover different altitude ranges. Measurements at lower altitudes require more spectral coverage, but less resolution, than those at higher altitudes. Standard 25-channel spectrometers are the primary source of information for measurements throughout the stratosphere. These have a bandwidth of 1300 MHz and resolution varying from 96 MHz at band edges to 6 MHz at band center. Individual channel positions and widths are given in Table III and illustrated in Fig. 7. Midband spectrometers, with lters having the same specications as the center eleven channels (numbered 8 to 18) in Table III, are used for additional measurements in the middle and upper stratosphere. Digital autocorrelator (DAC) spectrometers provide the ner spectral resolution for measurements of mesospheric temperature, H O, O , and",{"entities":[]}]
["Bourassa, A. E., Degenstein, D. A., Randel, W. J., Zawodny, J. M., Kyrl, E., McLinden, C. A., Sioris, C. E., and Roth, C. Z.: Trends in stratospheric ozone derived from merged SAGE II and Odin-OSIRIS satellite observations, Atmos. Chem. Phys., 14, 69836994, doi:10.5194/acp-14-6983-2014, 2014.",{"entities":[[188,192,"SPACECRAFT"],[193,199,"INSTRUMENT"]]}]
["Fig. 10. Time series of monthly mean tropospheric ozone VMR at the 4 cities shown in Fig. 7 (solid lines) along with corresponding time series averaged over 5 in latitude and longitude around the city grid cell (dashed lines)(3 month running average).",{"entities":[]}]
["Allen, D. R., J. L. Stanford, N. Nakamura, M. A. Lpez-Valverde, M. Lpez-Puertas, F. W. Taylor, and J. J. Remedios (2000), Antarctic polar descent and planetary wave activity observed in ISAMS CO from April to July 1992, Geophys. Res. Lett., 27(5), 665668, doi:10.1029/ 1999GL010888.",{"entities":[[186,191,"INSTRUMENT"]]}]
["Solomon, S., Garcia, R. R., Sherwood, R. F., and Wuebbles, D. J.: On the depletion of Antarctic ozone, Science, 321, 755758, https://doi.org/10.1038/321755a0, 1986.",{"entities":[]}]
["In Sect. 3.3, a sudden increase in temperature is observed over all three convective regions during the winter period in the pressure range 100 to 261 hPa. However, no striking change in WVMR or OLR values is observed during the time period in which temperature peaks appear. Furthermore, convection is absent over the selected regions during the boreal winter period (OLR > 260 W m2). Thus, convection may not be the source of such change in temperature. This is another interesting feature which is observed in this analysis and the reason of this increase must be inspected in details.",{"entities":[]}]
["gent track. On 10 January near the 65 N, 45 E and on 17 September near 65S, 225E, disagreement is greater than 10 K and may be associated with orthogonal LOS directions. ",{"entities":[]}]
["Zhang, J., Campbell, J.R., reid, J.S., Westphal, D.L., Baker, N.L., Hyer, E.J., 2011. Evaluating the impact of assimilating CALIOP-derived aerosol extinction profiles on a global mass transport model. Geophys. Res. Lett. 38, L14801. Zhang, J., Christopher, S.A., Remer, L., Kaufman, Y., 2005d. Satellite aerosol direct radiative forcing studies over cloud free oceans from Terra: (I) aerosol",{"entities":[[124,130,"INSTRUMENT"]]}]
["The tidal components have been extracted from the hourly winds of RMFR for 2002 and of both RMFR and KSS MR during 2010 winter, using the procedure discussed in section 2.1. The RMFR tidal amplitudes",{"entities":[[66,70,"INSTRUMENT"],[92,96,"INSTRUMENT"],[101,107,"INSTRUMENT"],[178,182,"INSTRUMENT"]]}]
["Stationary planetary waves are dominant in the Arctic stratosphere and strongly affect PSC areal extents. Their relative contribution to the total PSC coverage ratio is close to 100 % at most altitudes during the winter. Synoptic-scale waves and gravity waves act to decrease PSC coverage area, although their effects are very weak.",{"entities":[]}]
["Wheeler, A. J., Xu, X. H., Kulka, R., You, H. Y., Wallace, L., Mallach, G., Van Ryswyk, K., MacNeill, M., Kearney, J., Rasmussen, P. E., Dabek-Zlotorzynska, E., Wang, D., Poon, R., Williams, R., Stocco, C., Anastassopoulos, A., Miller, J. D., Dales, R., and Brook, J. R.: Windsor, Ontario Exposure Assessment Study: Design and Methods Validation of Personal, Indoor, and Outdoor Air Pollution Monitoring, J. Air Waste Manage., 61, 324338, doi:10.3155/1047-3289.61.3.324, 2011.",{"entities":[]}]
["Sze, N. D., Stratospheric fluorine: A comparison between theory and measurements, Geophys. Res. Lett., 5, 781-783, 1978. ",{"entities":[]}]
["In summary, OSIRIS ozone drifts very likely to higher values above 20 km. The drift is quite small up to 35 km and close to the 5 % signicance threshold. In the upper stratosphere the presence of a +(58) % decade1 drift is evident. The OSIRIS team has found that the drift in ozone may be caused by a positive drift in the altitude registration. Efforts are under way to correct for this in the next data release.",{"entities":[[12,18,"INSTRUMENT"],[236,242,"INSTRUMENT"]]}]
["For the 19841997 period, the data sets show relatively similar trend proles in all latitude bands, most with maximum negative trends in the upper stratosphere (top row Fig. 8 and Tables 810). In the southern mid-latitudes (Fig. 8a), SWOOSH, SAGE-OSIRIS, and both SAGE-GOMOS data",{"entities":[[246,252,"INSTRUMENT"],[268,273,"INSTRUMENT"]]}]
["Figure 1. Monthly sample count in 5 latitude bins for 16 instrumental sampling patterns. Gray regions denote regions of no measurements.",{"entities":[]}]
["[53] Changes from v4 to v5 MLS ozone data for the Arctic winter are typically not as large as those shown above for Antarctica (and the two data versions tend to track better).",{"entities":[[27,30,"INSTRUMENT"]]}]
["uncertainties are only estimates of random error and do not include any indications of overall accuracy.",{"entities":[]}]
["ever, few formal D&A methods have been applied in studies involving stratospheric ozone (see Gillett et al., 2011, for one exception to this). There is evidence that stratospheric ozone is transitioning from an era of widespread and readily detectable depletion (linked to changes in anthropogenic chlorouorocarbons) to an era characterized by early signs of recovery or healing (Solomon et al., 2016). Our motivation for this work is to determine whether formal D&A methods can provide a more condent and quantitative attribution of ozone depletion and recovery signals.",{"entities":[]}]
["The diagonal elements assumed for the radiance error co variance in clear sky and thin cirrus are 2 K between 80 and 316 hPa increasing to 5 K for tangent pressures greater than 464 hPa. These values are based on the mean residual of the radiance fit to the dry and wet continua rounded to 1 K. The radiance errors modeled in this way are expected to include contributions from limb pressure, temperature, contaminant species, and inadequacy of the continuum parameterization. Atmospheric scattering of the radiation is another consid eration. Noticeable scattering of 202/204 GHz radiation oc curs when cloud particle sizes exceed about 100 pm. The MLS radiative transfer forward model does not include scat tering and cannot model the brightness temperature depres sion seen in the lowest altitudes of the scan. Scattering can be detected by an unusually large negative value of observed ",{"entities":[[669,672,"INSTRUMENT"]]}]
["Siskind, D. E., Stevens, M. H., Englert, C. R., and Mlynczak, M. G.: Comparison of a photochemical model with observations of mesospheric hydroxyl and ozone, J. Geophys. Res., 118, 195 207, doi:10.1029/2012JD017971, 2013.",{"entities":[]}]
["[9] The Regional Acid Deposition Model Version 2 (RADM2) [Stockwell et al., 1990] is used as the chemistry driver, and the Modal Aerosol Dynamics Model for Europe (MADE/SORGAM) [Ackermann et al., 1998; Schell et al., 2001] is used as the aerosol driver. The anthropogenic emissions are provided by the 0.5 0.5 Reanalysis of the TROpospheric (RETRO) chemical composition over the past 40 years (http:/retro.enes.org/index.shtml) and the 1 1 Emission Database for Global Atmospheric Research (EDGAR) (http://www.mnp.nl/edgar/introduction). The initial and boundary conditions for chemical constituents and aerosols are obtained from the offline Model for Ozone and Related Chemical Tracers, version 4 (MOZART4) global chemical transport model with a 2.8 2.8 horizontal resolution [Emmons et al., 2010]. Comparing to Measurements Of Pollution In The Troposphere (MOPITT) and MODIS observations, the MOZART4 model outputs show a high bias for the columnintegrated CO over South America and a low bias for AOT in the nearby oceanic regions [Emmons et al., 2010]. Since the species O3 in the MOZART4 is unrealistic in the stratosphere and upper troposphere [Emmons et al., 2010], the species O3RAD (instead of O3) in the MOZART4 is used for the initial and boundary conditions of O3 in the WRFChem simulations. The O3RAD variable is relaxed to the ozone climatology in the stratosphere and to the MOZART calculated ozone in the troposphere. It should be noticed that the O3RAD in the upper troposphere is much higher than observations due to too strong stratospheric flux from reanalysis meteorological data sets [van Noije et al., 2004; Emmons et al., 2010].",{"entities":[[875,880,"INSTRUMENT"]]}]
["Figure 17. Mean and standard temperature deviations between the TEMPERA radiometer and the measurements from the different instruments and WACCM.",{"entities":[[64,71,"INSTRUMENT"]]}]
["Fig. 16. Mean SAGE III (dotted) and MLS (solid) water vapor mixing ratio proles for 8000 coincident events from mid-2004 through 2005.",{"entities":[[36,39,"INSTRUMENT"]]}]
["Gebhardt, C., Rozanov, A., Hommel, R., Weber, M., Bovensmann, H., Burrows, J. P., Degenstein, D., Froidevaux, L., and Thompson, A. M.: Stratospheric ozone trends and variability as seen by SCIAMACHY from 2002 to 2012, Atmos. Chem. Phys., 14, 831846, doi:10.5194/acp-14-831-2014, 2014.",{"entities":[[189,198,"INSTRUMENT"]]}]
["Recently, Hoyle (2011) presented convective tracer transport using different CTMs, among them the Oslo CTM2, using a different set of meteorological data than used here. To compare the differences in transport between Oslo CTM3 and CTM2, we do similar tracer studies with one tracer held constant at 1 ppm below 500 m altitude and above having a lifetime of 6 h (T6h), and a second 20-days lifetime tracer held constant at 1 ppt at the surface (T20d).",{"entities":[]}]
["Randall, C. E., V. L. Harvey, D. E. Siskind, J. France, P. F. Bernath, C. D. Boone, and K. A. Walker (2009), NOx descent in the Arctic middle",{"entities":[]}]
["Beatty, T.J. et al., 1989. Simultaneous radar and lidar observations of sporadic E and Na layers at Arecibo. Geophys. Res. Lett. 16 (9), 1019 1022.",{"entities":[]}]
["Friedl-Vallon, F., Gulde, T., Hase, F., Kleinert, A., Kulessa, T., Maucher, G., Neubert, T., Olschewski, F., Piesch, C., Preusse, P., Rongen, H., Sartorius, C., Schneider, H., Schnfeld, A., Tan, V., Bayer, N., Blank, J., Dapp, R., Ebersoldt, A., Fischer, H., Graf, F., Guggenmoser, T., Hpfner, M., Kaufmann, M., Kretschmer, E., Latzko, T., Nordmeyer, H., Oelhaf, H., Orphal, J., Riese, M., Schardt, G., Schillings, J., Sha, M. K., Suminska-Ebersoldt, O., and Ungermann, J.: Instrument concept of the imaging Fourier transform spectrometer GLORIA, Atmos. Meas. Tech., 7, 3565 3577, https://doi.org/10.5194/amt-7-3565-2014, 2014.",{"entities":[[539,545,"INSTRUMENT"]]}]
["In fact, our simulations show that the IAV in the UT/LS composition due to shifts in the locations of deep convection is larger than that due to variations in the mean ascent rate and depth/frequency of convection. The uniform tracer",{"entities":[]}]
["that the corresponding PV values change between different levels, due to the strong dependence of PV on altitude. At levels of the subtropical jet core around 360 K, the strong jet to the north of the monsoon masks the existence of the anticyclone transport barrier (Garny and Randel, 2013).",{"entities":[]}]
["Theoretical considerations by Rood and Douglas (1985) and Douglas et al. (1985) conrmed these basic results; however, they indicated that it is necessary to very carefully analyze the inuence of the dynamic terms on the ozone transport before the correlations between ozone and temperature could be attributed to any specic process.",{"entities":[]}]
["Citation: Wargan, K., S. Pawson, M. A. Olsen, J. C. Witte, A. R. Douglass, J. R. Ziemke, S. E. Strahan, and J. E. Nielsen (2015), The global structure of upper tropospherelower stratosphere ozone in GEOS-5: A multiyear assimilation of EOS Aura data, J. Geophys. Res. Atmos., 120, 20132036, doi:10.1002/ 2014JD022493.",{"entities":[[199,205,"INSTRUMENT"],[235,238,"INSTRUMENT"],[239,243,"SPACECRAFT"]]}]
["to which O3 levels were inuenced by local meteorological and chemical processes. The smaller the change is, the larger the inuence of the local contributions is.",{"entities":[]}]
["Tobo, Y., Iwasaka, Y., Zhang, D., et al., 2008. Summertime ozone valley over the Tibetan Plateau derived from ozonesondes and EP/TOMS data. Geophys. Res. Lett. 35, 16.",{"entities":[[126,128,"SPACECRAFT"],[129,133,"INSTRUMENT"]]}]
["Another cause of the disagreement, however, can be found in the trajectory technique. Because these tra jectory calculations are nondiffusive, they will preserve filamentary structures for longer periods of time than may be physically warranted. Hence, disagreements be tween new and old measurements can also be caused by model errors. ",{"entities":[]}]
["cm -s of the Ed model and 1.70x10 x cm -s of the Bx model. For the nighttime values the data sets dif fer more. Indeed, in Bordeaux the nighttime average is about 2.57+0.05x10 m cm -s, while the MLS Oa con centration is only 2.25+0.02x10 (cid:127) cm -a, giving agree ment only to within 13%. The two model nighttime values are slightly closer to the Bordeaux data (2.48 2.60)x10 (cid:127) cm -s than to the MLS data, since the agree ment with Bordeaux is within 3.5%, as compared to 9% with MLS. The night-to-day ratio (Figure 8) shows this discrepancy, since Oanoon/O3night ratio is close to 0.75 for MLS, while it is 0.65 for Bordeaux, close to the Ed model ratio of 0.63 but far from the Bx model ratio of 0.70. Once again, as previously mentioned, we must remark that the Ed model does show a daytime as maximum than both the Bx and the Bordeaux data, while MLS data do not show this early morning maximum. We will return to this in section 3.4 when we compare model calculations. ",{"entities":[[208,211,"INSTRUMENT"],[427,430,"INSTRUMENT"],[512,515,"INSTRUMENT"],[625,628,"INSTRUMENT"],[909,912,"INSTRUMENT"]]}]
["Seppl, A., P. T. Verronen, E. Kyrl, S. Hassinen, L. Backman, A. Hauchecorne, J. L. Bertaux, and D. Fussen (2004), Solar proton events of OctoberNovember 2003: Ozone depletion in the Northern Hemisphere polar winter as seen by GOMOS/Envisat, Geophys. Res. Lett., 31, L19107, doi:10.1029/2004GL021042.",{"entities":[[226,231,"INSTRUMENT"],[232,239,"SPACECRAFT"]]}]
["Fukao, S., et al., 1994. Seasonal variability of vertical eddy di7usivity in the middle atmosphere, 1. three-year observations by the middle and upper atmosphere radar. Journal of Geophysical Research 99, 18,97318,987.",{"entities":[]}]
["Tuck, A. F., S. J. Hovde, K. K. Kelly, M. J. Mahoney, M. H. Proffitt, E. C. Richard, and T. L. Thompson (2003), Exchange between the upper tropical troposphere and the lower stratosphere studied with aircraft observations, J. Geophys. Res., 108(D23), 4734, doi:10.1029/2003JD003399.",{"entities":[]}]
["3Norwegian Institute for Air Research, Kjeller, Norway. 4Jet Propulsion Laboratory, California Institute of Technology,",{"entities":[]}]
["3.1. The permittivity of pure ice Laboratory measurements of the imaginary part of the ice refractive index or dielectric permittivity have been made and published by a number of researchers at millimeter and sub-millimeter wavelengths; they are summarized in Figure 1 with comparison to the modied LH model (Equation (5)). The solid lines in Figure 1 show the imaginary ice dielectric permittivity computed by the modied LH model at different temperatures. The dotted lines are the original curves computed from the original LH formula (Equation (3)). The measured data are described in the following.",{"entities":[]}]
["Atmos. Chem. Phys., 15, 42154224, 2015 www.atmos-chem-phys.net/15/4215/2015/ doi:10.5194/acp-15-4215-2015 Author(s) 2015. CC Attribution 3.0 License.",{"entities":[]}]
["Chameides, W.: Tropospheric odd nitrogen and the atmospheric water vapor cycle, J. Geophys. Res., 84, 49894996, https://doi.org/10.1029/JC080i036p04989, 1975.",{"entities":[]}]
["Hanisco, T. F., et al. (2007), Observations of deep convective inuence on stratospheric water vapor and its isotopic composition, Geophys.",{"entities":[]}]
["[28] We used neural networks to correct for interinstrument biases and produce a consistent time series of HCl from 1991 to the present. Such an HCl time series is of use in estimating a time series of Cly .",{"entities":[]}]
["At 70 hPa (Fig. 7a), ozone and water vapour together can account for an annual cycle in temperature of about 2.8 0.3 K peak-to-peak, i.e. about 35 % of the observed annual cycle in temperature. The cancellation between the effects of ozone and water vapour on temperature is strongest at 90 hPa (Fig. 7b), with the combined amplitude being about 2.3 0.4 K peak-to-peak, i.e. again about 40 % of the observed annual cycle. At 100 hPa, the combined amplitude is about 1.50.4 K peak-to-peak or about 45 % of the observed annual cycle (not shown). Thus, while the estimated contribution of dynamical heating to the annual cycle in temperatures is substantially smaller than the observed annual cycle (Fig. 7ac), when the contributions from dynamical heating, ozone, and water vapour are combined the result is in remarkably good agreement with the observed annual cycle, both in amplitude and in phase.",{"entities":[]}]
["[2] L. L. Pan, C. R. Homeyer, S. Honomichl et al., Thunderstorms enhance tropospheric ozone by wrapping and shedding stratospheric air, Geophysical Research Letters, vol. 41, no. 22, pp. 77857790, 2015.",{"entities":[]}]
["Sato, K., M. Tsutsumi, T. Sato, T. Nakamura, A. Saito, Y. Tomikawa, K. Nishimura, M. Kohma, H. Yamagishi, and T. Yamanouchi (2014), Program",{"entities":[]}]
["Acknowledgements. This work is part of Jiali Luos PhD research, funded by the National Science Foundation of China (41705021, 41630421, and 41575038). The work was in part conducted at the National Center for Atmospheric Research, operated by the University Corporation for Atmospheric Research under sponsorship of the United States National Science Foundation. The IASI mission is a joint mission of EUMETSAT and the Centre National dEtudes Spatiales (CNES, France). We thank the ULB team (Daniel Hurtmans, Pierre Coheur) for the development of the FORLI-CO retrieval algorithm and Mijeong Park for helpful discussions. We also thank the three anonymous reviewers for their helpful comments and suggestions.",{"entities":[[367,371,"INSTRUMENT"]]}]
["Temperature data from the Microwave Limb Sounder (MLS) are used to estimate wavenumber and period of planetary waves during SSW. MLS is a limb scanning emission microwave radiometer on the NASA Aura satellite (Waters et al., 2006; Livesey et al., 2007). Aura was launched on July 15, 2004 on a sunsynchronous polar orbit at 705 km altitude with a 981 inclination. The MLS instrument scans the limb in the orbital plane which gives a global coverage from 821S to 821N on each orbit. The limb scans are made in the forward direction of the satellite. The useful height range of temperature data is approximately 897 km (3160.001 hPa) with a vertical resolution of 4 km in the stratosphere and 14 km at the mesopause determined by the full width at half maximum (FWHM) of the averaging kernels (Livesey et al., 2007). Here we use the temperature data from the level 2 version 2.2 data product. We removed poor data by screening methodologies described by Livesey et al. (2007). The geometric altitudes are approximately calculated from the pressure levels as follows: h 7 lnp=1000, where h is the altitude in km and p the pressure in hPa. Note that there is a difference between geometric and geopotential heights especially in the mesosphere. However, for studies of PWs and considering the altitudinal resolution of MLS in the mesosphere, this difference is not important.",{"entities":[[26,48,"INSTRUMENT"],[50,53,"INSTRUMENT"],[129,132,"INSTRUMENT"],[194,198,"SPACECRAFT"],[254,258,"SPACECRAFT"],[368,371,"INSTRUMENT"],[1319,1322,"INSTRUMENT"]]}]
["Manney, G. L. and R. W. Zurek, Interhemispheric compar ison of the development of the stratospheric polar vortex during fMl: a 3-dimensionM perspective for 1991-92, Geo phys. Res. Lett., 20, 1274-1278, 1993. ",{"entities":[]}]
["Figure 2. The TES degrees of freedom for the global survey of 3 4 July 2006. The blue data points show the number of pieces of information in the troposphere, while the red triangles show the dofs for the stratosphere. The discontinuities at 54S, 18S, 18N, and 54N are due to changes in the a priori constraint matrix used in the retrieval algorithm.",{"entities":[[14,17,"INSTRUMENT"]]}]
["ACE-FTS N2O proles were compared with measurements from Odin SMR (v2.1), Aura MLS (v2.2) and MIPAS (ESA, v4.62 and IMK-IAA, v3.0). The relative differences for these data sets are within 15 % between 6 and 30 km, except for MIPAS near 30 km where the differences are as large as 22.5 %. Between 30 and 60 km, the absolute differences for the satellite comparisons typically range from 2 to +1 ppbv. Relative differences are not given because the N2O VMR is quite small over this altitude range.",{"entities":[[0,7,"INSTRUMENT"],[56,60,"SPACECRAFT"],[73,77,"SPACECRAFT"],[78,81,"INSTRUMENT"],[93,98,"INSTRUMENT"],[224,229,"INSTRUMENT"]]}]
["Uppala, S., Kallberg, P., Simmons, A. J., Andrae, U., Bechtold, V. D. C., Fiorino, M., Gibson, J. K., Haseler, J., Hernandez, A., Kelly, G. A., Li, X., Onogi, K., Saarinen, S., Sokka, N., Allan, R. P., Andersson, E., Arpe, K., Balmaseda, M. A., Beljaars, A. C. M., van de Berg, L., Bidlot, J., Bormann, N., Caires, S., Chevallier, F., Dethof, A., Dragosavac, M., Fisher, M., Fuentes, M., Hagemann, S., Holm, E., Hoskins, B. J., Isaksen, L., Janssen, P. A. E. M., Jenne, R., McNally, A. P., Mahfouf, J. F., Morcrette, J. J., Rayner, N. A., Saunders, R. W., Simon, P., Sterl, A., Trenberth, K. E., Untch, A., Vasiljevic, D., Viterbo, P., and Woollen, J.: The ERA-40 re-analysis, Q. J. Roy. Meteor. Soc., 131, 29613012, 2005.",{"entities":[]}]
["According to the theory, buoyancy increases by the release of latent heat, and decreases when condensate loading (i.e., the weight of liquid or ice in a uid parcel) increases (see Eqs. (2.50)(2.53) of Houze, 2014). Lebo and Seinfeld (2011) state that the aerosol-induced effect is controlled by the balance between latent heating and the increase in condensed water aloft, each having opposing effects on buoyancy. Since changes in buoyancy can be positive or negative, depending upon specic situations in which latent heating or condensate perturbations are dominant, changes in cloud structure IWC likely could be positive or negative as AOD increases.",{"entities":[]}]
["Holm EV, Untch A, Simmons A, Saunders R, Bouttier F, Andersson E. 1999. Multivariate ozone assimilation in four-dimensional data assimilation. Pp 8994 in Proceedings of the SODA Workshop on Chemical Data Assimilation, 910 December 1998, KNMI, De Bilt, The Netherlands.",{"entities":[]}]
["The three-dimensional Chemical Transport Model (CTM) of the Laboratory for Atmospheres at the God dard Space Flight Center (GSFC) is used for this anal ysis. The model accurately simulates the detailed dis tribution of chemical tracers including ozone [Douglass et al., 1996] and carbon dioxide [$trahan et al., 1998]. The CTM is fully described by Douglass et al. [1996], but key parameters are repeated here. The three dimensional flux form semi-Lagrangian transport code [Lin and Rood, 1996] uses 46 level winds from the God dard Earth Observing System (GEOS) Assimilation System [Schubert et al., 1993]. The transport scheme is specifically designed for avoiding negative tracer values and for .conserving mass. Model resolution is 2 latitude by 2.5 longitude with 46 hybrid sigma-pressure levels in the vertical face between sigma and pressure occurs at -(cid:127)400 hPa, so that the region discussed in this work is all in pressure coordinates. The model pressure levels are -(cid:127)20 hPa (1 kin) apart near the tropical tropopause. The model is run in an off-line mode, so the tracer concentrations do not feedback to affect the winds and temperatures. Input winds and temperatures are interpolated to a 15 rain (900 s) time step from 6 hourly assimilation data. Explicit convective transport is not included. Vertical motion is calculated from the divergence of the hori zontal wind. Douglass et al. [1996] note two benefits of using an offline transport model over a full GCM with chemistry and feedbacks; the ability to have tem peratures more representative of the actual stratosphere than those in most GCMs and to appropriately simulate individual synoptic-scale perturbations for comparison to observations. Both of these benefits are critical to the simulations described in this work and distinguish this analysis from the GCM study of Mote et al. [1994]. For comparison of the simulation with observations, a variety of assimilated and observed data sets are used. These data sets are discussed more fully as each is in troduced; they range from global assimilation systems to individual ",{"entities":[[542,603,"INSTRUMENT"]]}]
["Table 1. Summary of OZORAM characteristics. The altitude range and altitude resolution depend on integration time. The gures given here correspond to a measurement interval of 1 h. For the difference in altitude range refer to Sect. 2.2. Data recorded until 2004 have been used in past publications (Sinnhuber et al., 1998; Klein et al., 2000, 2002) and are listed here for completeness.",{"entities":[[20,26,"INSTRUMENT"]]}]
["Olsen, M. A., M. R. Schoeberl, and A. R. Douglass (2004), Stratosphere-troposphere exchange of mass and ozone, J. Geophys. Res., 109, D24114,",{"entities":[]}]
["Figure 4. As in Figure 1 for the 1992-1993 northern hemisphere late-winter north-looking period. The slopes of the lines lead to estimates of the diabatic vertical velocity of 0.1 q0.2 mms -(cid:127) at 585 K and 0.1 q0.1 mms (cid:127) at 465 K. ",{"entities":[]}]
["Hoskins, B. J., M. E. Mcintyre, and A. W. Robertson, On the use and significance of isentropic potential vorticity maps, Q. J. R. Mctorol. Soc., 111,877-946, 1985. ",{"entities":[]}]
["Schreiner, W., Rocken, C., Sokolovskiy, S., Syndergaard, S., and Hunt, D.: Estimates of the precision of GPS radio occultations from the COSMIC/FORMOSAT-3 mission, Geophys. Res. Lett., 34, L04808, doi:10.1029/2006GL027557, 2007.",{"entities":[[137,143,"INSTRUMENT"],[144,154,"SPACECRAFT"]]}]
["CTM3 tropospheric mixing ratios are slightly higher south of 50 S, indicating slightly faster transport to southernmost latitudes. North of 50 S the differences are generally related to convective regions, where CTM2 has higher mixing ratio than CTM3 in the lower troposphere. This is mainly due to differences in entrainment and detrainment, and will be explained in Sect. 3.2.2.",{"entities":[]}]
["Alexander, M. J., Geller, M. A., McLandress, C., Polavarapu, S., Preusse, P., Sassi, F., Sato, K., Eckermann, S. D., Ern, M., Hertzog, A., Kawatani, Y., Pulido, M., Shaw, T. A., Sigmond, M., Vincent, R. A., and Watanabe, S.: Recent developments in gravity-wave effects in climate models and the global distribution of gravity-wave momentum ux from observations and models, Q. J. Roy. Meteor. Soc., 136, 11031124, doi:10.1002/qj.637, 2010.",{"entities":[]}]
["instrumental sampling biases is positive biases around 30N and S between 200 and 100 hPa. In these regions, the tropopause slopes downward with latitude, creating a strong horizontal O3 gradient on pressure surfaces. Instrumental sampling densities tend to increase modestly with latitude. As a result, within the latitude band that straddles the tropopause for a given pressure surface, instruments tend to sample the poleward side of the latitude band slightly more often than the equatorward side, leading to positive bias in the average O3 mixing ratio. This feature is apparent to different degrees in the UTLS sampling bias estimates for Aura MLS, HIRDLS, MIPAS, and TES. The sampling biases for ACE-FTS and OSIRIS are dominated by nonuniform month of year sampling issues as discussed above, and for SCIAMACHY, the nonuniformity of longitudinal sampling is seen to have an inuence between 20 and 70S, seemingly amplifying the magnitude of the tropopause-related bias at 30S.",{"entities":[[644,648,"SPACECRAFT"],[649,652,"INSTRUMENT"],[654,660,"INSTRUMENT"],[662,667,"INSTRUMENT"],[673,676,"INSTRUMENT"],[702,709,"INSTRUMENT"],[714,720,"INSTRUMENT"],[807,816,"INSTRUMENT"]]}]
["Brasseur, G., and C. Granier, Mount Pinatubo aerosols, chlorofluorocarbons and ozone depletion, Science, 257, 1239-1242, 1992. ",{"entities":[]}]
["tions in the channel that observed in the spectral window of 394620 nm (spectral channel 3). The instrument scanned the Earths limb from the surface up to 100 km with a 2.5 km vertical eld of view and a 3 km vertical sampling. The NO2 retrieval algorithm, detailed by Rozanov et al. (2005) and summarized by Bauer et al. (2012), uses limb-scattered radiances measured from 420 to 470 nm and solves the inverse problem using the DOAS technique and Tikhonov regularization (Tikhonov, 1963). In each prole, the spectra are normalized by the limb radiances nearest 43 km. The regularization matrix smooths the retrievals using an empirically determined height-dependent smoothing parameter, chosen in order to minimize physically unrealistic oscillations in proles while maximizing vertical resolution. The retrieval makes use of a forward model that takes into account absorption by O3 (simultaneously retrieved) and O2O2 and uses pressure and temperature proles from the European Centre for Medium-Range Weather Forecasts (ECMWF). The NO2 and O3 absorption cross sections were obtained from Bogumil et al. (1999). The algorithm retrieves NO2 proles between 10 and 40 km with a typical vertical resolution of 3 5 km, degrading to 10 km at the upper and lower retrieval altitude limits.",{"entities":[]}]
["Verhoelst, T., Granville, J., Hendrick, F., Khler, U., Lerot, C., Pommereau, J.-P., Redondas, A., Van Roozendael, M., and Lambert, J.-C.: Metrology of ground-based satellite validation: co-location mismatch and smoothing issues of total ozone comparisons, Atmos. Meas. Tech., 8, 50395062, doi:10.5194/amt-8-5039-2015, 2015.",{"entities":[]}]
["Sideband fraction Filter position Spectrometer nonlinearity Power supply interaction Hot reference standing waves Gain compression Radiometric calibration Field-of-View calibration",{"entities":[]}]
["Hertzog, A., Vial, F., Dornbrack, A., Eckermann, S. D., Knudsen, B. M., and Pommereau, J.-P.: In-situ observations of gravity waves and comparisons with numerical simulations during the SOLVE/THESEO 2000 campaign, J. Geophys. Res., 107(D20), 8292, doi:10.1029/2001JD001025, 2002.",{"entities":[]}]
["Figure 6. Same as Figure 3, but for comparison between MIPAS and Radiosonde temperatures. Also shown are the zonal mean differences derived for the MIPAS ascending (nighttime, bottom plots) orbit node. See color version of this figure in the HTML.",{"entities":[[55,60,"INSTRUMENT"],[148,153,"INSTRUMENT"]]}]
["Orsolini, Y. J., G. L. Manney, M. L. Santee, and C. E. Randall (2005), An upper stratospheric layer of enhanced HNO3 following exceptional solar storms, Geophys. Res. Lett., 32, L12S01, doi:10.1029/2004GL021588.",{"entities":[]}]
["[38] For the comparison to be entirely satisfactory, Dxrms should be the same as the combined error of the two data sets, while Dxmean should be much smaller. Because the CO",{"entities":[]}]
["Note that we do not attempt to derive temperatures from the meteor radar measured diffusion coefcients here, as for example, described in Holdsworth et al. (2006), as there are signicant limitations in this approach (see Lee et al., 2013; Younger et al., 2014). However, Lee et al. (2016) have described a new method of estimating temperatures near the mesopause region using meteor radar observations by calibrating their meteor radar against Aura MLS temperatures. This approach looks promising and has some similarities to the approach we describe below in Sect. 4.3.1.",{"entities":[[444,448,"SPACECRAFT"],[449,452,"INSTRUMENT"]]}]
["Figure 2. Time-longitude section of MLS temperature anomalies at 21.5 hPa for a 60-day period since 15 August 2005. The contour interval is 1.0 K with negative values shaded.",{"entities":[[36,39,"INSTRUMENT"]]}]
["Wohltmann, I., Lehmann, R., and Rex, M.: The Lagrangian chemistry and transport model ATLAS: simulation and validation of stratospheric chemistry and ozone loss in the winter 1999/2000, Geosci. Model Dev., 3, 585601, 2010.",{"entities":[]}]
["It is, however, evident that there is signicant correlation among the predictor variables as shown in Figure 3 (bottom panels), which is strongest between GPH and ISZA . The violation of the basic assumptions of normality, independence, and constant variance of the residuals and the predictor variables being measured without error leads to problems in least squares regression analysis (see Rawlings et al., 1998). Correlation among predictor variables, however, does not always result in collinearity. The inclusion or exclusion of a certain predictor for example H2O may increase or decrease collinearity.",{"entities":[]}]
["each band; we believe that such an accuracy gure is indeed achievable, with most of the uncertainty in this region arising from spectroscopic uncertainties or inconsistencies between the various bands. More complete error analyses, including the impact of any pointing knowledge uncertainties (currently believed to be a small contributing factor; see [11]) will be pursued later. The larger differences in the lower stratosphere,",{"entities":[]}]
["Boone, C. D., Walker, K. A., and Bernath, P. F.: Version 3 retrievals for the Atmospheric Chemistry Experiment Fourier Transform Spectrometer (ACE-FTS), in: The Atmospheric Chemistry Experiment ACE at 10: a Solar Occultation Anthology, A. Deepak Publishing, Hampton, Virginia, USA, 2013.",{"entities":[[78,141,"INSTRUMENT"],[143,150,"INSTRUMENT"]]}]
["(a) Contributions of trajectories from SCS, Figure 10. MON, TIB, and the combination of the three domains (CMB) to the mean simulated water vapor at 68 hPa during October through December 2004 and 2005 according to GMAO MERRA. (b) Same as Figure 10a but for ERA Interim. Error bars encompass twice the standard error of the mean.",{"entities":[]}]
["Froidevaux, L., et al. (2008), Validation of Aura Microwave Limb Sounder stratospheric ozone measurements, J. Geophys. Res., 113, D15S20,",{"entities":[[45,49,"SPACECRAFT"],[50,72,"INSTRUMENT"]]}]
["Barnard, J. C., J. D. Fast, G. Paredes-Miranda, W. P. Arnott, and A. Laskin (2010), Technical note: Evaluation of the WRF-Chem aerosol chemical to aerosol optical properties module using data from the MILAGRO campaign, Atmos. Chem. Phys., 10, 73257340, doi:10.5194/acp-107325-2010.",{"entities":[]}]
["situ measurements of aerosol microphysical properties in SEA, with none to our knowledge within Indochina itself. However, some information can be inferred from existing chemistry studies. Aged particles throughout SEA, including those from biomass burning in the MC, have high ionic fractions, which corroborate isolated reports of high hygroscopicity. Black",{"entities":[]}]
["Scarnato, B., Staehelin, J., Stbi, R., and Schill, H.: Long-term total ozone observations at Arosa (Switzerland) with Dobson and Brewer instruments (19882007), J. Geophys. Res., 115, D13306, doi:10.1029/2009JD011908, 2010.",{"entities":[]}]
["The WVMRs of FISH and DP30 show a very good agreement for most of the calibration conditions, except for the lowest and highest mixing ratios steps. Here, a dependence of the WVMR on the cell pressure can be seen (see Fig. 4), which is not considered in the linear FISH calibration (Eq. 1), and thus points to some deviations from the idealized measurement principle described above.",{"entities":[[13,17,"INSTRUMENT"],[22,26,"INSTRUMENT"],[265,269,"INSTRUMENT"]]}]
["vapor mixing ratios using the nominal UKMO temperatures and background aerosol conditions. Results for the extremes of 3 and 6 ppmv are shown in Plate 8a since they bracket the behavior of the predicted HNO3. Although the MLS data indicate water vapor mixing ratios on the order of 3 ppmv by mid-July, they do not support such extensive water va por loss in the late May/early June time frame [Santee et al., 1995]. Nevertheless, the top panel of Plate 8a shows that the agreement of both the NAT and NAD models with the data could be significantly better than that of the standard run if we have overestimated the water vapor abundance. In the case of water mixing ratios larger than those of our standard run, the bottom panel of Plate 8a shows even better agree ment between the data and the ternary solution model. ",{"entities":[[225,228,"INSTRUMENT"]]}]
["Crounse, J. D., Paulot, F., Kjaergaard, H. G., and Wennberg, P. O.: Peroxy radical isomerization in the oxidation of isoprene, Phys. Chem. Chem. Phys., 13, 1360713613,doi:10.1039/c1cp21330j, 2011.",{"entities":[]}]
["Rind, D., E.-W. Chiou, W. Chu, S. Oltmans, J. Lerner, J. Larsen, M.P. McCormick, and L. McMaster, Overview of the Stratospheric Aero sol and Gas Experiment lI water vapor observations: Method, val idation, and data characteristics, J. Geophys. Res., 98, 4835-4856, 1993. ",{"entities":[]}]
["[10] The Aura satellite is in a sun-synchronous orbit. The time when the MLS field of view scans the latitudes and longitudes covering TMF is often between 2100 2130 UT with a small variation from day to day. To be accurate, the SZA is used as the primary criterion to match the TMF data points with the MLS measurements for any given day.",{"entities":[[9,13,"SPACECRAFT"],[73,76,"INSTRUMENT"],[305,308,"INSTRUMENT"]]}]
["Additional ltering that has been applied to the Aura MLS WV data to remove low-biased data in the UTLS is described in Appendix A. This ltering is motivated by the dry bias present in the MLS data, as described in the next section. This appendix describes a new algorithm for screening out Aura MLS WV proles in the UTLS that are dry biased and contain unphysical oscillations. The motivation for Appendix A is that rather than simply remove all MLS data below a predetermined pressure level, it is desirable to retain as much of the Aura MLS data in the UTLS region (e.g., at lower latitudes) that are thought to be unaffected by the dry bias. The screening procedure described in Appendix A has been applied to all Aura MLS WV data stored in SWOOSH, so no action on the part of SWOOSH users is required.",{"entities":[[48,52,"SPACECRAFT"],[53,56,"INSTRUMENT"],[188,191,"INSTRUMENT"],[446,449,"INSTRUMENT"],[534,538,"SPACECRAFT"],[539,542,"INSTRUMENT"],[717,721,"SPACECRAFT"],[722,725,"INSTRUMENT"]]}]
["[38] The rising Aura/MLS N2O mixing ratios are interpreted as ascent of the vortex in the vortex descent study above. The Odin/SMR N2O mixing ratios do on the other hand not rise in late February so the Odin/SMR vortex descent study does not indicate any vortex ascent in that period. Furthermore, the cross-isentropic transport rates from the SLIMCAT model corroborate the Odin/SMR descent rates and do not indicate vortex ascent in late February. Since the vertical gradients of Aura/MLS N2O mixing ratios inside the vortex drop more rapidly as a function of altitude after mid-February than in the early winter, the rising N2O mixing ratios in the lower vortex could possibly be explained by increased exchange of air masses across the vortex border at low altitudes. For this explanation to be true, the cross vortex border exchange would however have had to be stronger than seen in the passive DIAMOND model and it is furthermore not corroborated by the Odin/SMR N2O measurements. As mentioned in section 2.2 of this manuscript, significant discrepancies exist between the version 1.5 Aura/MLS N2O data used in this study and the more recently",{"entities":[[16,20,"SPACECRAFT"],[21,24,"INSTRUMENT"],[122,126,"SPACECRAFT"],[203,207,"SPACECRAFT"],[208,211,"INSTRUMENT"],[374,378,"SPACECRAFT"],[379,382,"INSTRUMENT"],[481,485,"SPACECRAFT"],[486,489,"INSTRUMENT"],[960,964,"SPACECRAFT"],[965,968,"INSTRUMENT"],[1091,1095,"SPACECRAFT"],[1096,1099,"INSTRUMENT"]]}]
["The aerosol rst indirect effect is related to the radiative forcing caused in the change of cloud particle size and hence the modication of the single-scattering properties of a cloud due to the presence of larger aerosol loadings. For the aerosol effect on De, we follow the approach presented by Jiang et al. (2011), in which an analytical formula was obtained to describe the variations of De with IWC and AOD for several regions with distinct characteristics of De-IWC-AOD. Using MLS IWC measurements at 215 hPa to indicate convective strength (CONV) and AOD measurements from MODIS to denote aerosol loading, as well as using a least-squares tting, an empirical formula for ice cloud effective radius re (De/2) as a function of CONV and AOD was derived to approximately capture the observed relationships among these three parameters as follows: re = AOD[1exp(CONV/)]exp(CONV), (1)",{"entities":[[484,487,"INSTRUMENT"],[581,586,"INSTRUMENT"]]}]
["greater chemical loss. Chemical loss and transport of ozone cases. The largest ozone decrease rates observed in the Arctic are typically one-half to two-thirds those in the Antarctic at the equivalent time of year, and chemical ozone destruction stops much earlier in the season in the Arctic. As was mentioned previously, because Antarctic loss is larger and replenishment by descent is smaller, transport masks very little of the ozone loss seen there [Manney et al., 1995a]. Mtiller et al. [1996] compared HALOE CH 4 and ozone relationships for the Arctic vortex in early and late winter during the winters of 1991-1992, 1992-1993, 1993-1994, and 1994-1995. Their results, which are broadly consistent with those from the MLS, show evidence of chemical ozone loss in each year. ",{"entities":[[751,754,"INSTRUMENT"]]}]
["Fig. 1. DC-8 individual ight tracks in INTEX-B colored by altitude. 17 science ights were performed from 4 March to 15 May, 2008.",{"entities":[]}]
["We have presented a novel approach to identify and account for data artefacts that remain in multiple ozone composites of satellite observations. These artefacts are one of",{"entities":[]}]
["Figure 17 shows the mean relative ozone bias of MACC ozone with respect to ozonesondes and MOZAIC data averaged between 200 and 1000 hPa averaged over the period from January 2003 to December 2010. With respect to ozonesondes, the reanalysis biases are within (510) % in the NH and over the Antarctic, but larger negative biases are found in the tropics. With respect to MOZAIC data, the reanalysis has mainly positive biases of less than 10 % over Europe and North and South America, and negative biases of up to 10 % over Africa. Larger positive biases with respect to MOZAIC are found over east Asia. Larger biases over east Asia were also seen for CO data (Fig. 6), suggesting that either the horizontal resolution is not high enough to reproduce the high values seen over polluted airports, that there could be problems with the vertical transport, or that the",{"entities":[]}]
["5.2.3 Albedo Above 30 km, agreement between OSIRIS and the validation data sets was found to be related to the OSIRIS albedo, which measures apparent upwelling and is obtained by tting the absolute value of the 740 nm modelled limb radiance at 40 km by adjusting the albedo with a forward model (Bourassa et al., 2007). The strongest bias was observed at 42.5 km, and is shown in Fig. 10 for comparisons with MLS and GOMOS. At all latitudes, OSIRIS ozone measurements are larger for higher albedo. At these altitudes, UV wavelengths are used in the ozone retrievals, so very little limb-scattered sunlight would originate from the lower altitudes, suggesting that this bias is not caused by errors in the radiative transfer due to poor estimates of albedo. Therefore,",{"entities":[[45,51,"INSTRUMENT"],[112,118,"INSTRUMENT"],[411,414,"INSTRUMENT"],[419,424,"INSTRUMENT"],[444,450,"INSTRUMENT"]]}]
["Randel, W. J., and M. Park (2006), Deep convective inuence on the Asian summer monsoon anticyclone and associated tracer variability",{"entities":[]}]
["Dvortsov, V. L. and Solomon, S.: Response of the stratospheric temperatures and ozone to past and future increases in stratospheric humidity, J. Geophys. Res., 106, 75057514, doi:10.1029/2000JD900637, 2001.",{"entities":[]}]
["Roswintiarti, O., Raman, S., 2003. Three-dimensional simulations of the mean air transport during the 1997 forest fires in Kalimantan, Indonesia using a mesoscale numerical model. Pure Appl. Geophys. 160, 429438.",{"entities":[]}]
["Li, J. M., Yi, Y. H., Minnis, P., Huang, J. P., Yan, H. R., Ma, Y. J., Wang, W. C., and Ayers, J. K.: Radiative eect dierences between multi-layered and single-layer clouds derived from CERES, CALIPSO, and CloudSat data, J. Quant. Spectrosc. Ra., 112, 361375, 2011. 24930",{"entities":[[206,214,"SPACECRAFT"]]}]
["Figure 3. SH polar projections of mean MIL frequency for (top row) WACCM, (middle row) SABER, and (bottom row) MLS in the SH between (left column) 15 May and 15 July and (right column) 15 July and 15 September. Thick black contours are as in Figure 2. Anticyclones are present at least 10% of the time equatorward of the thick gray contour during 15 May to 15 July and 30% of the time during 15 July to 15 September. Thick black contour encompasses regions where the vortex is present 30% during 15 May to 15 July and 60% during 15 July to 15 September.",{"entities":[[111,114,"INSTRUMENT"]]}]
["Randel, W. J., Wu, F., & Rios, W. R. (2003). Thermal variability of the tropical tropopause region derived from GPS/MET observations. Journal of",{"entities":[]}]
["Figure 3. Same as Figure 2 for zonal wavenumber 2, at 21 hPa. The contour spacing is 0.001 K 2 for temperature and 0.000025 ppmv 2 for ozone. ",{"entities":[]}]
["particles are generally smaller than the mean free path (at a typical stratosphere condition with 50 hPa and 200 K, the mean free path is about 0.8 mm). For those particles which have large Kn, B is simplied as B 5 1.666Kn and the fall velocity equation can be p written in kinetic limit: p=2makT",{"entities":[]}]
["Jensen, E. J., and G. E. Thomas (1994), Numerical simulations of the effects of gravity waves on noctilucent clouds, J. Geophys. Res., 99(D2),",{"entities":[]}]
["cycle of O3 we follow the procedure described in Randel et al. (2007) where ozone and temperature observations of the seven SHADOZ station closest to the equator (see subpanel in Fig. 2 for their geographical position) are considered, but instead of pressure we use potential temperature as the vertical coordinate. In particular, p-related observations of each station are transformed to -levels using the measured temperatures and then averaged over all seven stations for each level. A clear annual cycle of O3/hO3i, with the highest values of O3/hO3i in late summer and early fall in the range between 370 and 430 K, can be diagnosed from the SHADOZ data. The lowest values appear approximately 4 5 months earlier. The pattern of this very pronounced cycle is similar to the analysis on the p-levels (Chae and Sherwood, 2007; Randel et al., 2007), although O3/hO3i is signicantly smaller with peak anomaly 0.5 versus 0.3 during summer by using p and as the vertical coordinate, respectively (Fueglistaler et al., 2009a; Konopka et al., 2009). As discussed by Konopka et al. (2009), a signicant part of the variability of O3/hO3i on the p-levels is a seasonal adiabatic process (as the p-levels move relative to the -levels during the year because of the seasonal cycle in temperature) that can be removed by using potential temperature as the vertical coordinate. The seasonality of O3 above the level of zero clear sky radiation (Q=0 level around 360 K as discussed in Gettelman et al., 2004) was recently explained as a consequence of the annual cycle in the tropical upwelling with the strongest and weakest upwelling in winter and summer, respectively (Randel et al., 2007). Randel et al. (2007) showed that the",{"entities":[[124,130,"INSTRUMENT"],[647,653,"INSTRUMENT"]]}]
["Mueller, R., J.-U. Grooss, C. Lemmen, D. Heinze, M. Dameris, and G. Bodeker (2008), Simple measures of ozone depletion in the polar",{"entities":[]}]
["([(cid:127)MX iv for directions within solid angle flMX) 2. The variation of GM(v, O, rk) is small across the range of frequencies accepted by any one radiometer ",{"entities":[]}]
["El Amraoui, L., Peuch, V.-H., Ricaud, P., Massart, S., Semane, N., Teyss`edre, H., Cariolle, D., and Karcher, F.: Ozone loss in the 20022003 Arctic vortex deduced from the assimilation of Odin/SMR O3 and N2O measurements: N2O as a dynamical tracer, Q. J. Roy. Meteor. Soc., 134, 217228, 2008a.",{"entities":[[188,192,"SPACECRAFT"],[193,196,"INSTRUMENT"]]}]
["Trends in Os calculated using the GSFC 2D interac tive model are asymmetric and qualitatively similar to the SBUV trends, unlike other models that have been used in profile Os trend calculations. The model trend is due to large dynamical differences between the north ern and southern hemispheres, particularly in the fall and winter. These dynamical differences result in lower values of CH4 and NO in the southern hemisphere high latitudes than in the northern hemisphere. Since these constituents play an important role in Cly partition",{"entities":[[110,114,"INSTRUMENT"]]}]
["FIG. 6. Seasonal cycle estimates of CH 4 in January, April, July, and October. Contour interval is 0.1 ppmv.",{"entities":[]}]
["In the previous sections, we examined the sensitivity of OH and its interhemispheric gradient to model biases in tropospheric and stratospheric ozone, water vapor, and NH NOx emissions. We nd that water vapor has the largest impact on global mean OH, while NH NOx emissions have the largest impact on the NH/SH ratio. However, none of these biases individually explains the 20 % reduction in NH OH that would remove the interhemispheric asymmetry and which Figs. 4 and 5 suggest are necessary to remove most of the CO bias. We note, however, that the OH parameterization simulations do not account for all the chemical feedbacks between ozone, methane, OH, and other species, and may underestimate the sensitivity of the full chemistry simulation to some of these biases.",{"entities":[]}]
["Randall, C. E., Lumpe, J. D., Bevilacqua, R. M., Hoppel, K. W., Fromm, M. D., Salawitch, R. J., Swartz, W. H., Lloyd, S. A., Kyro, E., von der Gathen, P., Claude, H., Davies, J., DeBacker, H., Dier, H., Molyneux, M. J., and Sancho, J.: Reconstruction of three-dimensional ozone elds using POAM III during the SOLVE, J. Geophys. Res., 107(D20), 8299, doi:10.1029/2001JD000471, 2002.",{"entities":[]}]
["Despite the linear regime summarised in Sect. 5, there will be always regions where a single linearisation point will be unlikely to be adequate (like the polar vortex). In such cases, several linearisation points may be used to cover more atmospheric variability and the one that minimises a 2 function (i.e. the square of the residuals between the measurements and the expected spectrum for the linearly adjusted prole) is selected as the solution.",{"entities":[]}]
["aData version, vertical range, vertical resolution, and references for the six limb-viewing instruments used in this study.",{"entities":[]}]
["Thayer, J. P., M. Rapp, A. J. Gerrard, E. Gudmundsson, and T. J. Kane (2003), Gravity wave influences on Arctic mesospheric clouds as determined by a Rayleigh lidar at Sondrestrom, Greenland, J. Geophys. Res., 108(D8), 8449, doi:10.1029/2002JD002363.",{"entities":[[150,164,"INSTRUMENT"]]}]
["[22] The production of geophysical data (Level 2 data) from the calibrated observations of atmospheric limb radiances (Level 1 data) involves the Level 2 retrieval algorithms [Livesey et al., 2006]. These employ an optimal estimation method [Rodgers, 1976, 2000] applied to the problem of a nonlinear weighted least squares minimization of a cost function involving the fit to the observed Level 1 radiance signals with regularization provided by a priori constraints. The MLS forward model [Read et al., 2006; Schwartz et al., 2006] takes into account the physics of the radiative transfer process and instrument specific parameters to calculate radiance estimates given a particular atmospheric state. The Level-2 processor invokes an inverse model that uses the forward model and a priori constraints in an iterative scheme, starting from an initial guess atmospheric state (obtained from a climatology), to determine the optimal atmospheric state.",{"entities":[[473,476,"INSTRUMENT"]]}]
["As indicated in the previous section, most emission data sets are provided without any information on uncertainties on the data used for quantifying the emissions. Several sources of uncertainties have been identied, which will be summarised in this section.",{"entities":[]}]
["-3-2-10123CTM DIAL (ppmv)2009/09/012009/10/012009/11/01Date (yyyy/mm/dd)-3.0-2.5-2.0-1.5-1.0-0.50.0sPV (10-4 s-1)08 hPa, mean diff. = -0.39 ppmv, rms diff. = 1.08 ppmv for 51.6 S / 70.3 WT. Sugita et al.: Comparison of ozone proles in Ro Gallegos, Argentina",{"entities":[]}]
["the level where physical processes may lead to irreversible STE event. In addition, there is also a tropopause break from 10 km altitude down to 6.5 km altitude (Fig. 3a) stretching a line along the cyclonic-shear side of the eastern jet streak (Fig. 3b) from northwest of Spain to East of Denmark. The tropopause break and the parallel PV band stretching on 315 K (Fig. 2) are dynamical signatures of on-going upper-level frontogenesis and tropopause folding (Keyser and Shapiro, 1986). This kind of fold is generally a consequence of an ageostrophic circulation near a jet streak, where the air from the lower stratosphere is intruded into the troposphere. In our case, we notice horizontal winds in excess of 50 m/s, showing an upper-level jet streak with southwest",{"entities":[]}]
["The brightness temperature as measured in what is referred to as the line direction, Tb,line, from the ground can be divided into the microwave background radiation T0 plus the water vapor emission line originating in the middle atmosphere T ma b,z , scaled with the airmass factor for the line measurement, and the continuum emission in the troposphere approximated as mean tropospheric temperature Teff. For the part of the radiation originating above the tropopause, namely T0 and T ma b,z , one needs to consider the attenuation in the troposphere: Tb,line = (T ma line+ T0)e b,z Ama Here A is the airmass factor of the specied atmospheric layer (ma = middle atmosphere, trop = troposphere) for the specied measurement and z is the tropospheric opacity in the zenith direction.",{"entities":[]}]
["Plieninger, J., von Clarmann, T., Stiller, G. P., Grabowski, U., Glatthor, N., Kellmann, S., Linden, A., Haenel, F., Kiefer, M., Hpfner, M., Laeng, A., and Lossow, S.: Methane and nitrous oxide retrievals from MIPAS-ENVISAT, Atmos. Meas. Tech., 8, 46574670, doi:10.5194/amt-8-4657-2015, 2015.",{"entities":[[210,215,"INSTRUMENT"],[216,223,"SPACECRAFT"]]}]
["equivalent radiance (NER) at 9.6 mm (channel 4) is on the 1, which is equivalent to the order of 10 blackbody temperature of 88 K. Relying on the low noise level in the infrared range SABER is able to achieve a vertical resolution less than 2 km without degradation as altitude increases. These advantages place SABER in somewhat of a unique position with a high potential for studying the MLT region.",{"entities":[]}]
["Clerbaux, C., Coheur, P.-F., Hurtmans, D., Barret, B., Carleer, M., Semeniuk, K., McConnell, J. C., Boone, C., and Bernath, P.: Carbon monoxide distribution from the ACE-FTS solar occultation measurements, Geophys. Res. Lett., 32, L16S01, doi:10.1029/2005GL022394, 2005.",{"entities":[[166,173,"INSTRUMENT"]]}]
["Fig. 2. Time series of the annual 4th highest MDA8 O3 and the 3-year running average of the 4th highest MDA8 O3 at GBNP for 1994 to 2013. Dotted lines indicate the current federal standard (75 ppb) and the upper end of the range for a potential revised federal standard (70 ppb).",{"entities":[]}]
["Temperature is a key parameter for dynamical, chemical and radiative processes in the atmosphere. There exist several techniques to measure atmospheric temperature proles like radiosonde (e.g. Luers, 1997; Rufeux and Joss, 2003), FTIR (Fourier transform infrared, e.g. Smith et al., 1999; Feltz et al., 2003), lidar (e.g. Evans et al., 1997; Alpers et al.,",{"entities":[[230,234,"INSTRUMENT"]]}]
["Masarik, J. and Beer, J.: Simulation of particle uxes and cosmogenic nuclide production in the Earths atmosphere, J. Geophys. Res., 104, 1209912111, 1999.",{"entities":[]}]
["1Chemical Sciences Division, Earth System Research Laboratory, National Oceanic and Atmospheric Administration, Boulder, Colorado, USA. 2Cooperative Institute for Research In Environmental Sciences,",{"entities":[]}]
["Cunnold, D. M., W. P. Chu, R. A. Barnes, M. P. McCormick, and R. E. Veiga, 1989: Validation of SAGE II ozone measurements. J. Geophys. Res., 94, 84478460.",{"entities":[]}]
["UARS Microwave Limb Sounder, Gravity Wave Processes, NATOASI Series I.' Global Environmental Change, 50, 103-120, 1997. ",{"entities":[[0,4,"SPACECRAFT"],[5,27,"INSTRUMENT"]]}]
["Figure 7. Schematic diagram showing the variability of CPH (brown colour line) and COH (magenta colour line) with respect to the centre of cyclone. Spiral bands of convective towers reaching as high as COH are shown with blue colour lines. Light blue (red) colour up (down) side arrow shows the up drafts (downdrafts/subsidence). Thickness of the arrows indicates the intensity.",{"entities":[]}]
["Shindell, D. T.: Inhomogeneous forcing and transient climate sensitivity, Nat. Clim. Chang., 4, 274277, doi:10.1038/nclimate2136, 2014.",{"entities":[]}]
["Chemical ozone loss in the Arctic stratosphere has been observed since 1989. Since then, cold winters are prone to large chemical ozone loss due to the still high amounts of ozone depleting substances in the atmosphere (Rex et al., 2004). However, because of large planetary wave activity, the polar vortex breaks up or dissipates early in most Arctic winters (WMO, 2011; Harris et al., 2010; Kuttippurath et al., 2010b; Manney et al., 2003). Therefore, the vortex persistence has been comparatively shorter and the associated ozone loss smaller in the Arctic as compared to the Antarctic (WMO, 2011; Solomon et al., 2007). The longest vortex persistence in the Arctic was found in 1996/1997, in which the wave activity was considerably suppressed, and therefore the vortex was sustained until early May (Lef`evre et al., 1998; Coy et al., 1997). Nevertheless, the ozone loss in 1996/1997 was lower than that of other cold winters such as 1994/1995, 1999/2000, and 2004/2005 due to relatively higher temperatures in DecemberFebruary 1996/1997, when chlorine activation plays a key role in determining the magnitude of ozone loss (Manney et al., 2003; Santee et al., 1997). In contrast, very low temperatures were observed in MarchApril due",{"entities":[]}]
["[19] C. P. Davis, C. Emde, and R. S. Harwood, A 3-D polarized reversed Monte Carlo radiative transfer model for millimeter and submillimeter passive remote sensing in cloudy atmospheres, IEEE Trans. Geosci. Remote Sens., vol. 43, no. 5, pp. 10961101, May 2005.",{"entities":[]}]
["Portmann, R. W., S. Solomon, R. R. Garcia, L. W. Thomason, L. R. Poole, and M. P. McCormick (1996), Role of aerosol variations in anthropogenic",{"entities":[]}]
["[32] Daily vortex averages of Aura/MLS and Odin/SMR N2O measurements are plotted in Figure 5. The dashed lines are 5-day running averages of the N2O mixing ratios north of 75 equivalent latitude and the solid lines are 5-day",{"entities":[[30,34,"SPACECRAFT"],[35,38,"INSTRUMENT"],[43,47,"SPACECRAFT"],[48,51,"INSTRUMENT"]]}]
["Fig. 1. The 2007 seasonal behavior of NLC occurrences and NLC brightness as well as the MLS mesopause temperature (black line) and frost point temperature (blue line) at 0.0046 mb or 85 km at 601N. Nights with NLCs are indicated by scaled points with appropriate digits of the NLC brightness. The red curve is a Gaussian t in the leastsquare sense to the MLS mesopause temperature. Gaps are due to the Aura data voids on day 194, 219 and 220 of year. (For interpretation of the references to color in this gure legend, the reader is referred to the web version of this article.)",{"entities":[[88,91,"INSTRUMENT"],[355,358,"INSTRUMENT"]]}]
["Independent measurements of atmospheric state were obtained from the Aura MLS (Microwave Limb Sounder) instrument [Waters et al., 2006]. Launched in 2004, the MLS instrument aboard the Aura EOS satellite measures microwave emissions across ve bands ranging from 118 GHz to 2.5 THz. Data products include the density of a number of dierent atmospheric species, neutral temperature, and geopotential height. Observations are made every 165 km along the orbital track, producing vertical proles of each parameter gridded on xed pressure surfaces.",{"entities":[[69,73,"SPACECRAFT"],[74,77,"INSTRUMENT"],[79,101,"INSTRUMENT"],[159,162,"INSTRUMENT"],[185,189,"SPACECRAFT"]]}]
["Manney, G. L., et al. (2008b), The high Arctic in extreme winters: Vortex, temperature, and MLS and ACE-FTS trace gas evolution, Atmos. Chem. Phys., 8, 505522, doi:10.5194/acp-8-505-2008.",{"entities":[[92,95,"INSTRUMENT"],[100,107,"INSTRUMENT"]]}]
["Weiss, A. K., Staehelin, J., Appenzeller, C., and Harris, N. R. P.: Chemical and dynamical contributions to ozone prole trends of the Payerne (Switzerland) balloon soundings, J. Geophys. Res., 106, 2268522694, doi:10.1029/2000JD000106, 2001.",{"entities":[]}]
["Siskind, D. E., Marsh, D. R., Mlynczak, M. G., Martin-Torres, F. J., and Russell III, J. M.: Decreases in atomic hydrogen over the summer pole: Evidence for dehydration from polar mesospheric clouds?, Geophys. Res. Lett., 35, L13809, https://doi.org/10.1029/2008GL033742, 2008.",{"entities":[]}]
["[11] Figure 1 shows the zonal and temporal dependence of the NLC occurrence rate (Figure 1a) and the temperature fluctuations at 85 km (Figure 1b) for the period June 9 June 24, 2005. This period was chosen, because it is one with the most obvious quasi 5-day wave signature in the NLC occurrence rates. Both panels show a clear westward propagating (i.e., occurrence rate maxima and minima propagate from east to west) wave signature of wavenumber 1 (i.e., the zonal variation on a given day corresponds to one 2p cycle of a sine wave) and a period of about 5 days. Therefore, the signature corresponds to the quasi 5-day wave, or the symmetric Rossby (1, 1) normal mode oscillation. Temperature minima coincide with maxima in the NLC occurrence rate and vice versa. The excellent agreement between the wave signature in the NLC occurrence rates and the mesopause temperature fluctuations suggests that the quasi 5-day wave signature in the temperature field is the main driver for the NLC wave signature. The temperature fluctuations show maximum quasi 5-day wave amplitudes of about 3 K. Note, that the actual temperature amplitude at 85 km may differ from the observed one, because of the rather poor vertical resolution of the MLS temperature profiles near the mesopause (12 15 km). [12] Figure 2 shows the temporal evolution of the NLC occurrence rates for the [120, 90] longitude bin and illustrates that the quasi 5-day wave can have a very large impact on the NLC occurrence rates for a certain geographical location. This is in line with the ground-based NLC sightings reported by Kirkwood and Stebel [2003].",{"entities":[[1233,1236,"INSTRUMENT"]]}]
["It is clear that large scale horizontal transport can be resolved by limb observations. The transport into and out of the $H vortex exatnined here at 22 hPa seems somewhat limited, especially dur ing the early and midwinter when the vortex is strong. This is not the case near 10 hPa in both hemispheres where there is clear evi dence of frequent poleward transport of 0-3.. Although the horizon tal transport does not appear to result in an exchange of air from outside the vortex to inside at this level, the descent implied at these upper altitudes could result in ozone-rich air entering the vortex at a lower level since the edge and shape of the vortex ",{"entities":[]}]
[" 4. Soden, B. J. et al. Quantifying Climate Feedbacks Using Radiative Kernels. Journal of Climate 21, 35043520, https://doi.",{"entities":[]}]
["There is indication of intraseasonal variability in cloud ozone over the eastern and western Pacic Ocean regions and also over Central America. In the western Pacic the intraseasonal variability originates largely from the 12-month MaddenJulian oscillation. In the eastern Pacic the largest variability is interannual and originates from ENSO and associated changes in SST/convection. In the eastern Pacic the highest cloud ozone occurs during La Nia (suppressed convection over the region) with lowest cloud ozone during El Nio (enhanced convection).",{"entities":[]}]
["Acknowledgements The authors thank Marc Allaart (KNMI) for his contributions to the calculation of PSC formation temperatures, Holger Vomel at CIRES, University of Colorado (USA) for providing a webpage with a list of available water vapor saturation pressure formula (http:// cires.colorado.edu/,voemel/vp.html), and Ronald van der A, Peter van Velthoven and Peter Siegmund (KNM) for their comments and suggestions. This work was partly financed by the Netherlands Space Office as part of the SCIAVISIE project.",{"entities":[]}]
["Yuan, T., Remer, L.A., Pickering, K.E., Yu, H., 2011. Observational evidence of aerosol enhancement of lightning activity and convective invigoration. Geophys. Res. Lett. 38, L04701. http://dx.doi.org/10.1029/2010GL046052. Yumimoto, K., Uno, I., Sugimoto, N., Shimizu, A., Liu, Z., Winker, D.M., 2008. Adjoint inversion modeling of Asian dust emission using lidar observations. Atmos. Chem. Phys. 8, 28692884.",{"entities":[]}]
["Komhyr, W. D., Barnes, R. A., Brothers, G. B., Lathrop, J. A., and Opperman, D. P.: Electrochemical concentration cell ozonesonde performance evaluation during STOIC 1989, J. Geophys. Res., 100, 92319244, doi:10.1029/94JD02175, 1995.",{"entities":[]}]
["Wang, D. Y., von Clarmann, T., Fischer, H., Funke, B., Gil-Lopez, S., Glatthor, N., Grabowski, U., Hopfner, M., Kaufmann, M., Kellmann, S., Kiefer, M., Koukouli, M. E., Linden, A., LopezPuertas, M., Mengistu Tsidu, G., Milz, M., Steck, T., Stiller, G. P., Simmons, A. J., Dethof, A., Swinbank, R., Marquardt, C., Jiang, J. H., Romans, L. J., Wickert, J., Schmidt, T., Russell III, J., and Remsberg, E.: Validation of stratospheric temperatures measured by Michelson Interferometer for Passive Atmospheric Sounding MIPAS on Envisat, J. Geophys. Res., 110, D08301, doi:10.1029/2004JD005342, 2005.",{"entities":[[456,513,"INSTRUMENT"],[514,519,"INSTRUMENT"],[523,530,"SPACECRAFT"]]}]
["Waters, J. W., L. Froidevaux, W. G. Read, G. L. Manney, L. S. Elson, D. A. Flower, R. F. Jarnot, and R..S. Har wood, Stratospheric C10 and ozone from the Microwave Limb Sounder on the Upper Atmosphere Research Satel lite, Nature, 36, 597-602, 1993. ",{"entities":[[164,187,"INSTRUMENT"],[195,231,"SPACECRAFT"]]}]
["Figure 13. Vertical ozone prole (in nmol mol1) at De Bilt (Netherlands) for (a) June, (b) December 2008. In (a) results for all EMAC and the two COSMO/MESSy instances are shown, while (b) shows the results for EMAC and CM50. The standard deviation of the temporal mean is indicated by the error bars for the observations and by the shaded area for the simulation data.",{"entities":[]}]
["Subtracting the monthly averages of the 00:0024:00 UTC temperatures from the 01:0003:00 and 10:0012:00 UTC temperatures gave the estimated biases in Aura daily means due to only sampling during some hours of the day and are given in Fig. 3. The gure shows that by judging by the measurement windows, Aura underestimates the daily mean (00:0024:00 UTC) more during winter that during spring and summer. Note the higher standard deviations in spring and summer compared to winter.",{"entities":[[149,153,"SPACECRAFT"],[300,304,"SPACECRAFT"]]}]
["Edinburgh University Jet Propulsion Laboratory Herlot-Watt University Lockheed Palo Alto Research Laboratory ",{"entities":[]}]
["[40] Unlike the 10 January case, the global distribution of the vertical flux of zonal momentum on 20 January (Figure 5b) reveals less vertical coherence among the regions of westward propagating GWs at various levels.",{"entities":[]}]
["Convective overshoots that penetrate the tropopause in both the tropics and the extra-tropics have the potential to increase the humidity of the stratosphere, including the stratospheric overworld, through the rapid sublimation of convectively lofted ice. This has been demonstrated in both modeling (tropics: Grosvenor et al. [2007], Jensen et al. [2007], and Dessler et al. [2007]; extra-tropics: Wang [2003], Dessler and Sherwood [2004], Le and Gallus [2012], and Homeyer et al. [2017]) and observational studies (tropics: Kley et al. [1982], Corti et al. [2008], de Reus et al. [2009], Khaykin et al. [2009], Iwasaki et al. [2010], Sayres et al. [2010], Sargent et al. [2014], and Khaykin et al. [2016]; extra-tropics: Poulida et al. [1996], Hegglin et al.",{"entities":[]}]
["longitude. Moreover, the aircraft flight path on the way back to the airport in Houston, Texas, was retracing the outbound flight path, but at higher altitude (17 km versus 14 km for the outbound flight). In addition, the reduction of the ozone column above the aircraft following the change of latitude (from 30 to 20 degrees north) can be seen in the first half of the time series shown in Figure 7. Profile sensitivity errors",{"entities":[]}]
["line), C1ONO2 (dotted line), HOC1 (dashed line), and C1 (thick line) calculated by the 0-D model at 10 ((cid:127)31 kin), 4.6 ((cid:127)36 kin), 2.1 ((cid:127)42 kin), 1 ((cid:127)47 kin), (cid:127)nd 0.46 hPa ((cid:127)53 km). The night-to-day increase in C10 is attributed below 40 km to the diurnal variation in C1ONO2 and between 40 and 50 km to variations in HOC1. At 40 km the diurnal variations of both C1ONO2 and HOC1 influence the C10 evolution. Above 50 km the C10 variation is no longer anticorrelated with HOC1 since it shows a night-to-day decrease that is mainly due to a repartitioning within the C1Ox family proportional to the 0/03 ratio as mentioned in section 2. ",{"entities":[]}]
["Gandrud, B. W., J. E. Dye, D. Baumgardner, G. V. Ferry, M. Loewenstein, K. R. Chan, L. Sanford, B. Gary, and K. Kelly, The January 30, 1989, Arctic polar stratospheric clouds (PSC) event: Evidence for a mechanism of dehydration, Geophys. Res. Lett., 17, 457-460, 1990. ",{"entities":[]}]
["Fig. 2b shows the Taylor diagram for selected radiation/aerosol/ cloud related variables from four simulations in 2006. There are some outliers including the AOD from two simulations (i.e., NCSU and EPA) due to negative correlation and LWP from EC due to a large NSD. All simulations show a good agreement for SWDN, LWDN, ORL, and PWV. Simulations generally overestimate the amplitude of variability for SWDN (except NCSU), LWDN, and CF and underestimate it for most of other variables. Correlation is excellent for SWDN, LWDN, OLR, and PWV (typically > 0.9) and good for CF and LWP (typically between 0.6 and 0.9), which is consistent with Figs. 3 and 4. The negative correlation for AOD is mainly caused by the large overpredictions over western U.S. and slightly underprediction over eastern U.S. Overall, the results show the high uncertainties in simulating many cloud related variables and further model improvement for the related physical/chemical treatments (e.g., aerosol activation scheme and aqueous-phase chemistry scheme) is warranted.",{"entities":[]}]
["Dewan, E.M., 1997. Saturated-cascade similitude theory of gravity wave spectra. Journal of Geophysical Research 102, 29,79929, 817.",{"entities":[]}]
["We also nd another maximum ozone loss near 650 K, as in Grooss and Muller (2007), but our estimate for 1 February to 10 March is 0.4 ppmv, which is clearly smaller than their value (around 0.6 ppmv for the 1 February to 10 March period).",{"entities":[]}]
["Figure 3. Mean NO proles (in ppbv) for coincident diurnally scaled ACE-FTS (black) and HALOE (left panel) and MIPAS IMKIAA (right panel) measurements and corresponding measurement standard deviations (dashed lines). Coincidence criteria for HALOE comparisons are within 3 h and 500 km and are within 3 h and 100 km for MIPAS IMK-IAA comparisons.",{"entities":[[67,74,"INSTRUMENT"],[110,115,"INSTRUMENT"],[319,324,"INSTRUMENT"]]}]
["[23] Acknowledgments. We thank GMAO and the HWV science team for data sets used. Work at the Jet Propulsion Laboratory, California Institute of Technology, was carried out under contract with NASA.",{"entities":[]}]
["significant impact on the time-average variance. Those using the flat spectrum bear a very close resemblance to those computed using the time mean winds and temper atures and so are not shown. In the narrow spectrum case (thin solid lines in Figure 18), however, use of the time-varying atmosphere has significantly modified the longitudinal structure of the time-mean variances, re sulting in a reduction of the peak magnitude of the fil tered variances by at least a factor of 2. Thus temporal variations become increasingly more important as the range of ground-based phase speeds in the source spec trum is reduced. These results indicate that the day to-day variability of the background winds can have a large impact on the time-average variances in the case of the narrow anisotropic spectrum, a result that has pre viously been shown by Dunkerton and Butchart [1984] in the context of stationary gravity wave propagation during sudden stratospheric warmings. ",{"entities":[]}]
["La Nina and El Ninoinduced variabilities of ozone in the tropical lower atmosphere during 19702001, Geophys. Res. Lett., 30, 1142, doi:10.1029/2002GL016387, 2003.",{"entities":[]}]
["Gates WL, Mitchell JFB, Boer GJ, Cubasch U, Meleshko VP (1992) Climate modelling, climate prediction and model validation. In: Houghton JT, Callander BA, Varney SK (eds) Climate change 1992, the supplementary report to the IPCC scientific assessment. Cambridge University Press, Cambridge, pp 97134",{"entities":[]}]
["decreasing to 10 pptv at southern remote areas. (The Thornton et al. (1999) data are also included in the comparison of Fig. 7.)",{"entities":[]}]
["Di Girolamo, L., Bond, T.C., Bramer, D., Diner, D.J., Fettinger, F., Kahn, R.A., Martonchik, J.V., Ramana, M.V., Ramanathan, V., Rasch, P.J., 2004. Analysis of Multi-angle Imaging SpectroRadiometer (MISR) aerosol optical depths over greater India during winter 20012004. Geophysical Review Letters 31, L23115. http://dx.doi.org/10.1029/2004GL021273.",{"entities":[[160,197,"INSTRUMENT"],[199,203,"INSTRUMENT"]]}]
["Detrainment in monsoon outow plumes and subsequent long-range transport in the upper troposphere export pollution from the ASM region to the Middle East, Africa, the Mediterranean, Europe, North America, and the Southern Hemisphere [e.g., Scheeren et al., 2003; Jiang et al., 2007; Liang et al., 2007; Barret et al., 2008; Lawrence and Lelieveld, 2010; Rogal et al., 2010; Vogel et al., 2014, 2015, 2016; Mller et al., 2016; Rauthe-Schch et al., 2016]. The monsoon circulation plays a key role in stratosphere-troposphere exchange during northern summer [e.g., Chen, 1995; Dunkerton, 1995; Dethof et al., 1999]. Observational and modeling studies have shown that deep convection in the ASM/Tibetan Plateau region enables a transport pipeline complementary to the slow large-scale ascent that takes place in the tropics throughout the year, allowing air to eectively bypass the lowest temperatures of the tropical tropopause layer and rapidly enter the lower stratosphere in the subtropics during boreal summer directly and/or via advection in the monsoon circulation [e.g., Gettelman et al., 2004; Fu et al., 2006; James et al., 2008; Park et al., 2009; Randel et al., 2010; Wright et al., 2011; Chen et al., 2012; Bergman et al., 2012, 2013; Heath and Fuelberg, 2014; Orbe et al., 2015; Garny and Randel, 2016; Pan et al., 2016].",{"entities":[]}]
["[36] The strong temperature sensitivity of ozone may be explained as follows: The ozone mixing ratio is determined by the ozone production (O2 photolysis, insensitive to temperature) and ozone destruction. The main catalytic ozone destruction cycles are NOx and ClOx cycles. Minor contributions to ozone loss are added by the Chapman cycle",{"entities":[]}]
["In the present work, the partial OH column (between the surface and 21.5 hPa) is estimated on the basis of OH profiles derived from the GEOS (Goddard Earth Observing System)-Chem model (for the troposphere) and the Harvard 2-D model (between the tropopause and 21.5 hPa). We also compare with observations from various other instruments and calculations from a constrained photochemical box model [Canty et al., 2006] to estimate the uncertainty of the lower atmospheric OH from the model, particularly for the lower stratosphere. The estimated lower atmospheric OH is combined with the MLS OH data to obtain a total OH column at latitudes and longitudes similar to TMF, which is then compared to the TMF OH column data at the Aura satellite overpass time. The overall agreement and the agreement at various seasons and solar zenith angles",{"entities":[[136,140,"INSTRUMENT"],[587,590,"INSTRUMENT"],[727,731,"SPACECRAFT"]]}]
["Clarisse, L., Coheur, P. F., Prata, A. J., Hurtmans, D., Razavi, A., Phulpin, T., Hadji-Lazaro, J., and Clerbaux, C.: Tracking and quantifying volcanic SO2 with IASI, the September 2007 eruption at Jebel at Tair, Atmos. Chem. Phys., 8, 77237734, doi:10.5194/acp-8-7723-2008, 2008.",{"entities":[[161,165,"INSTRUMENT"]]}]
["[4] Several studies have compared subsets of the analyses listed above or compared one or more of them with other local temperature data sets. Manney et al. [1996b] found that NCEP temperatures were consistently closer to radiosonde temperatures and lower than those from the Met Office during the 1991/1992 and 1994/1995 Arctic winters. Knudsen [1996], Knudsen et al. [1996], and Pullen and Jones [1997] found similar warm biases in ECMWF and Met Office temperatures with respect to sondes and other balloon observations in several Arctic winters. Pawson et al. [1999] compared temperatures from the FUB data with those derived from geopotential heights from the TIROS Operational Vertical Sounding (TOVS) system, and showed that the FUB temperatures were generally lower, but with large dispersion around the mean difference. Manney and Sabutis [2000] showed that Met Office minimum temperatures were lower than those from NCEP in January 2000. Davies et al. [2003] found that in cold regions Met Office temperatures were lower than ECMWF temperatures in January 2000 but higher in February 2000. They also showed that chemical transport model (CTM) runs driven with ECMWF and Met Office fields produced significantly different patterns of denitrification, chlorine activation, and ozone loss.",{"entities":[[664,669,"SPACECRAFT"]]}]
["Newell, R. E., Y. Zhu, E. V. Browell, W. G. Read, and J. W. Waters, Walker circulation and tropical upper tropospheric water vapor, J. Geophys. Res., 101, 1961-1974, 1996. ",{"entities":[]}]
["Ballard, J., et al, Absolute absorption coefficients of C1ONO2 infrared bands at stratospheric temperatures, J. Geophys. Res., 93, 1659-1665, 1988. ",{"entities":[]}]
["Soeva, V. F., Kyrola, E., Verronen, P. T., Seppala, A., Tamminen, J., Marsh, D. R., Smith, A. K., Bertaux, J.-L., Hauchecorne, A., Dalaudier, F., Fussen, D., Vanhellemont, F., Fanton dAndon, O., Barrot, G., Guirlet, M., Fehr, T., and Saavedra, L.: Spatiotemporal observations of the tertiary ozone maximum, Atmos. Chem. Phys., 9, 44394445, doi:10.5194/acp-9-4439-2009, 2009.",{"entities":[]}]
["level clouds in this set of models shows a wide range of values: TOMCAT/pTOMCAT underestimate the percentage of gridboxes with clouds tops above 12 km (or 13 km for West Africa), OSLOCTM2/FRSGCUCI, UMUKCA UCAM nud and, to a smaller extent, WRF tend to overestimate the percentage of gridboxes having clouds with tops above 13 14 km, while pTOMCAT tropical, UM UCAM highres and CATT-BRAMS show cloud heights which are either slightly lower, or within the observed range, depending on the region. Although the vertical distribution of clouds for the higher resolution models is generally closer to the observed range, horizontal resolution is not a major factor in determining the vertical distribution of clouds: in fact pTOMCAT tropical has a cloud distribution which is closer to observations compared to TOMCAT/pTOMCAT, despite having the same horizontal resolution and the same dynamical elds driving the large scale ow. A more detailed analysis of convection parameterisation in the TOMCAT-based models can be found in Feng et al. (2010). We now assess the ability of models to reproduce the relative strength of convection in the Maritime Continent compared to West Africa: OSLOCTM2/FRSGCUCI, UMUKCA UCAM nud, WRF and UM UCAM highres all show generally larger fractions of high-level clouds for the Maritime Continent compared to West Africa, which is consistent with observations; all the TOMCAT-based models however, show larger fractions of high clouds for West Africa compared to the Maritime Continent.",{"entities":[]}]
["This study also has some implications on CRE evaluation. Studies have shown that CRE in the UT region also aected the cross-tropopause mass transport of atmospheric constituents (Corti et al., 2006). Cloud inhomogeneity within satellite footprint has been treated with sophisticated schemes by some satellite observational teams",{"entities":[]}]
["Kero, A., C.F. Enell, A. J. Kavanagh, J. Vierinen, I. Virtanen, and E. Turunen (2008), Could negative ion production explain the polar mesosphere winter echo (PMWE) modulation in active HF heating experiments?, Geophys. Res. Lett., 35, L23102, doi:10.1029/2008GL035798.",{"entities":[]}]
["The large vertical gradient in the MLS CO VMR from mesosphere to the low stratosphere (Fig. 2) makes it a good tracer to study the downward transport of CO induced by the solar forcings in the lower thermosphere. The chemical balance between the source and the sink leads to the strong vertical gradient of CO mixing ratios with relatively abundant CO in the upper mesosphere, but with extremely small VMRs in the lower stratosphere. Horizontal CO gradients reect the transport in the winter polar vortex. The descent in the polar region brings high CO from the mesosphere to the lower stratosphere. The long CO lifetime over 30 days in the polar stratosphere and mesosphere during winter helps to maintain this vertical and horizontal gradients.",{"entities":[[35,38,"INSTRUMENT"]]}]
["Livesey, N. J., Filipiak, M. J., Froidevaux, L., Read, W. G., Lambert, A., Santee, M. L., Jiang, J. H., Pumphrey, H. C., Waters, J. W., Coeld, R. E., Cuddy, D. T., Daffer, W. H., Drouin, B. J., Fuller, R. A., Jarnot, R. F., Jiang, Y. B., Knosp, B. W., Li, Q. B., Perun, V. S., Schwartz, M. J., Snyder, W. V., Stek, P. C., Thurstans, R. P., Wagner, P. A., Avery, M., Browell, E. V., Cammas, J.-P., Christensen, L. E., Diskin, G. S., Gao, R.-S., Jost, H.-J., Loewenstein, M., Lopez, J. D., Nedelec, P., Osterman, G. B., Sachse, G. W., and Webster, C. R.: Validation of Aura Microwave Limb Sounder O3 and CO observations in the upper troposphere and lower stratosphere, J. Geophys. Res., 113, D15S02, doi:10.1029/2007JD008805, 2008.",{"entities":[[567,571,"SPACECRAFT"],[572,594,"INSTRUMENT"]]}]
["Fig. 5. Top row: Northern Hemisphere VITA (left) and Replay (right) N2O at 850 K on 28 March 2005. Red (blue) indicates high (low) N2O values. Overlaid are two EOS Aura MLS orbit tracks for this day. Second row: MLS (black), Replay (blue), and VITA (red) N2O at 850 K along the selected orbit tracks. The proles start at the locations of the orbit labels in the upper left panel. Replay and VITA data are interpolated linearly in space and time to the MLS data points.",{"entities":[[164,168,"SPACECRAFT"],[169,172,"INSTRUMENT"],[212,215,"INSTRUMENT"],[452,455,"INSTRUMENT"]]}]
["[32] The winter mean PSC and s2 distribution of Figure 4 suggested the important role of mesoscale orographic gravity waves in determining PSC occurrence and composition at 60S70S throughout winter. The detailed analysis of periods of large wave activity in the Hvmoller diagrams and the downstream effects for individual events (Figures 5 and 8) confirmed this initial interpretation.",{"entities":[]}]
["Livesey, N. J., Read, W. G., Froidevaux, L., Lambert, A., Manney, G. L., Pumphrey, H. C., Santee, M. L., Schwartz, M. J., Wang, S., Coeld, R. E., Cuddy, D. T., Fuller, R. A., Jarnot, R. F., Jiang, J. H., Knosp, B. W., Stek, P. C., Wagner, P. A., and Wu, D. L.: Earth Observing System (EOS), Aura Microwave Limb Sounder (MLS), Version 3.3 Level 2 data quality and description document, D-33509, Jet Propulsion Laboratory, California Institute of Technology, Pasadena, California, 2011.",{"entities":[[291,295,"SPACECRAFT"],[296,318,"SPACECRAFT"],[320,323,"SPACECRAFT"]]}]
["Indonesian forest re events in 1994 and in 1997 as revealed by ground-based observations, Geophys. Res. Lett., 26(16), 24172420, doi:10.1029/1999GL900117.",{"entities":[]}]
["Further investigation showed that the large amplitude oscillations in the temperatures estimated from the Kunming meteor radar could be due to the temperature sensitivity of the slope estimated from the meteor radar data. Therefore, an improved method was developed to calibrate the larger uctuations obtained by the temperature gradient model. The temperatures calibrated with this method were more consistent with the SABER observations, and there was a signicant improvement in the accuracy of the estimated temperatures.",{"entities":[]}]
["Klein, S. A., B. J. Soden, and N.C. Lau (1999), Remote sea surface variations during ENSO: Evidence for a tropical atmospheric bridge, J. Clim., 12, 917932, doi:10.1175/1520-0442(1999)012<0917:RSSTVD>2.0. CO;2.",{"entities":[]}]
["However, deep convection may play an opposite role on SWV by modulating the thermal structure of the TTL (e.g., Randel et al., 2015). In the tropics, a localized temperature minimum is frequently observed over active convection (e.g., Gettelman et al., 2002; Holloway & Neelin, 2007; Johnson & Kriete, 1982). Particularly, the cold anomaly is generally observed near the cold-point tropopause (CPT) as a thin layer lasting a couple of weeks in boreal winter (Kim & Son, 2012; Paulik & Birner, 2012). This tropopause-level cooling may dehydrate the lower stratosphere in a signicant manner because it makes the CPT colder, and air parcels generally travel for a couple of weeks near the CPT before they enter into the lower stratosphere (e.g., Fueglistaler et al., 2005).",{"entities":[]}]
["Yang, K., Liu, X., Krotkov, N.A., Krueger, A.J., Carn, S.A., 2009b. Estimating the altitude of volcanic sulfur dioxide plumes from space-borne hyper-spectral UV measurements. Geophys. Res. Lett. 36, L10803. http://dx.doi.org/10.1029/2009GL038025.",{"entities":[]}]
["GBNP is well situated to routinely intercept free tropospheric air through impaction and entrainment associated with convective mixing or subsidence (cf. Gustin et al., 2015; Lin et al., 2012b). Cooper et al. (2012) reported an increase of 0.41 0.27 ppb yr1 in free tropospheric O3 over western North America in the spring from 1995 to 2011, which is consistent with the range of increase detected at GBNP in May. Parrish et al. (2009) reported a signicant increase in marine boundary layer (MBL) inow at sites on the Pacic coast of the US of 0.46 0.13 ppb yr1 (spring); 0.43 0.17 ppb yr1 (winter); and 0.24 0.16 ppb yr1 (summer) for a period spanning from 1989 to",{"entities":[]}]
["Santer BD, Bonfils C, Painter JF, Zelinka MD, Mears C, Solomon S, Schmidt GA, Fyfe JC, Cole JN, Nazarenko L (2014) Volcanic contribution to decadal changes in tropospheric temperature. Nat Geosci 7:185189",{"entities":[]}]
["Kley, D., Russell III, J. M., and Phillips, C. (Eds.): Assessment of Upper Tropospheric and Stratospheric Water Vapour, SPARC Rep. 2, WMO, TD 1043, World Meteorol. Organ., Geneva, Switzerland, 2000.",{"entities":[]}]
["1 Bordeaux Observatory, CNRS/INSU, Floirac, France. 2NASA Langley Research Center, Hampton, Virginia. 3 Jet Propulsion Laboratory, Pasadena, California. 4 Edinburgh University, Edinburgh, Scotland, U.K. SHeriot-Watt University, Edinburgh, Scotland, U.K. ",{"entities":[]}]
["Figure 15 shows the vortex-averaged mixing ratios at 46 hPa simulated by Polar SWIFT in the Northern Hemisphere at the end of the winter compared to the mixing ratios obtained from MLS ozone data. Note that the date used in the plot differs for every year, since the date of the breakup of the polar vortex is different in every year. The dates are given in Table 4. Figure 16 shows the same for the Southern Hemisphere and on 1 October. Both the magnitude and the interannual variability of the MLS measurements are reproduced well by the Polar SWIFT model runs in the Northern Hemisphere. The interannual variability is larger and reproduced better in the Northern Hemisphere than in the Southern Hemisphere.",{"entities":[[181,184,"INSTRUMENT"],[496,499,"INSTRUMENT"]]}]
["3% higher than the SAGE II coincident values. This small level of disagreement is reduced to 1% when MLS v2.2 data are used [Froidevaux et al., 2007]. MLS partial stratospheric column ozone values for v2.2 have also been compared to ozonesonde partial columns [Jiang et al., 2007], and although there is often variability of order 5 to 10%, the mean differences are 1 to 2%, with MLS values on the high side, on average. The estimated accuracy for MLS column abundances down to near 215 hPa is 4%, based on sensitivity tests and simulated retrievals. Livesey et al. [2008] have provided additional validation results for MLS ozone (and CO), with a focus on the upper troposphere and lower stratosphere, including lidar and in situ data taken during the AVE aircraft campaigns.",{"entities":[[101,104,"INSTRUMENT"],[151,154,"INSTRUMENT"],[380,383,"INSTRUMENT"],[448,451,"INSTRUMENT"],[621,624,"INSTRUMENT"]]}]
["Finally, as mentioned, we note that two major SSW occurred in the Northern Hemisphere in January 2006 and January 2009 (a major warming being dened as a reversal of the winds at 10 hPa at a latitude of 60). It is known that major SSW can have dampening effect on planetary waves at high latitudes. In particular, it has been observed that planetarywave amplitudes can be suppressed after major SSW events (e.g. Alexander and Shepherd, 2010). If the ducting hypothesis is correct, then the MLT summer-time wave originates in the winter hemisphere and so any changes in wave amplitude due to major SSW may be reected in reduced wave amplitudes in the summer MLT of the opposite hemisphere. To see if such effects are present in our analysis, we examined the UKMO stratospheric winds and temperatures at 10 hPa to characterise the two major SSW. Figure 12 presents contours of these zonal-mean winds and temperatures for the six northern-hemisphere winters observed (the Southern Hemisphere is not considered because no major SSW occurred there during the observations). From the gure it can",{"entities":[]}]
["Holdsworth, D. A., R. J. Morris, D. J. Murphy, I. M. Reid, G. B. Burns, and W. J. R. French (2006), Antarctic mesospheric temperature estimation using the Davis MST radar, J. Geophys. Res., 111, D05108, doi:10.1029/2005JD006589. Jarvis, M. J. (2001), Bridging the atmospheric divide, Science,",{"entities":[]}]
["Schiller, C., A. Wahner, U. Platt, H-P. Dom, J. Callies, and D. H. Ehhalt, Near UV atmospheric absorption measure ments of column abundances during Airborne Arctic Stratospheric Expedition, January-February 1989, 2, OC10 observations, Geophys. Res. Lett., 17, 501-504, 1990. ",{"entities":[]}]
["1. formation of a dehydrated layer at 19.524 km (450 550 K) with a reduction in water mixing ratios of up to 1.6 ppmv;",{"entities":[]}]
["[57] The AquaVIT results are consistent with the lowwater calibrations performed at Harvard that constrained any potential Lyman-a instrument artifact to <0.2 ppmv in air. Furthermore, we have shown that these laboratory results are applicable to flight data. Therefore, we conclude that the differences observed in flight between HWV and CFH over the past two decades and most recently with JLH [see Gensch et al., 2008] are not explained by the AquaVIT results.",{"entities":[[331,334,"INSTRUMENT"],[339,342,"INSTRUMENT"],[392,395,"INSTRUMENT"]]}]
["Figure 2. The dotted lines show the averaging kernel func tions for the airborne radiometer. The solid line is the rel ative a priori contribution according to equation (2). The original 5-km spaced retrieval grid was converted to pres sure coordinates for this plot. ",{"entities":[]}]
["Santee, M. L., G. L. Manney, L. Froidevaux, W. G. Read, and J. W. Waters (1999), Six years of UARS Microwave Limb Sounder HNO3 observations: Seasonal, interhemispheric, and interannual variations in the lower stratosphere, J. Geophys. Res., 104, 8225 8246.",{"entities":[[94,98,"SPACECRAFT"],[99,121,"INSTRUMENT"]]}]
["retrieved C10 to be up to 0.4 ppbv too large at 46 hPa and 0.2 ppbv too large at 100 hPa; complete removal of N20 can cause the retrieved C10 to be up to 0.2 ppbv too small at 100 hPa. The effect of N20 is <0.05 ppbv at 46 hPa and above for this simulation. The effect of corn",{"entities":[]}]
["Lee, S.-H., Wilson, J. C., Reeves, J. M., and Laeur, B. G.: Aerosol size distributions from 4 to 2000 nm measured in the upper troposphere and lower stratosphere, in: European Aerosol Conference, Madrid, Spain, 2003.",{"entities":[]}]
["0.10.20.30.40.50.60.70.80.911.11.2Amplitude / ppmv90756045301501530456075901011001011021255443AmplitudeLatitude / degreePressure / hPaJFMAMJJASONDTime of first maximum / month90756045301501530456075901011001011021255443PhaseLatitude / degree(a)(b)S. Lossow et al.: Comparison of H2O variability",{"entities":[]}]
["is possible that the decline in HNO3 concentrations after 1993 apparent in the maps for mid-January (Plate 1) and, to a lesser extent, for mid-July (Plate 2) is related to the slow decay of the Pinatubo aerosol cloud. This is consistent with the gradual decreases in ground-based HNO3 column amounts that have been observed a t' southern midlatitudes [Koike et al., 1994] and northern low (cid:127)[David et al., 1994] and high [Slusser et al., 1998] latitudes as the stratosphere recov ered from Pinatubo-induced HNO3 enhancement. Kurner et al. [1996] also reported a distinct decreasing trend in UARS CLAES HNO3 abundances, especially in the south ern hemisphere, and attributed it to the diminishing influence of heterogeneous reactions as the Pinatubo aerosols settled out. We too have observed decreasing trends in the MLS HNO3 abundances at midlatitudes, a more detailed analysis of which is beyond the scope of this paper. The trends in the MLS midlatitude data have also been discussed by Randel et al. [ 1999]. ",{"entities":[[612,616,"SPACECRAFT"],[617,622,"INSTRUMENT"],[841,844,"INSTRUMENT"],[968,971,"INSTRUMENT"]]}]
["attempt to reduce this aerosol-PSC ambiguity by incorporating the additional 1064-nm backscatter coefcient and polarization information available in the CALIPSO data products. For instance, inclusion of the 532-nm perpendicular backscatter coefcient data may allow us to detect opticallythin layers of NAT particles.",{"entities":[]}]
["of material pulled off the vortex and intrusions into the vor tex in the lower stratosphere [e.g., Waugh et al., 1994; Plumb et al., 1994]; other studies show small horizontal scale struc tures along the vortex edge and in the anticyclone throughout the stratosphere [e.g., Pierce and Fairlie, 1993;Sutton et al., 1994]. Orsolini [ 1995] and Schoeberl and Newman [ 1995] demonstrated how small horizontal scale structures such as these which tilt with height may lead to lamination in ide alized tracers. Orsolini et al. [1997] used ozone from the Upper Atmosphere Research Satellite (UARS) Microwave Limb Sounder (MLS) to initialize high-resolution transport calculations and demonstrated that advection by the large scale wind field gave rise to small vertical scale structures re sembling those observed in a lidar ozone profile in the lower stratosphere through the process of filamentation. ",{"entities":[[567,602,"SPACECRAFT"],[604,608,"SPACECRAFT"],[610,633,"INSTRUMENT"],[635,638,"INSTRUMENT"]]}]
["Signicant perturbations were observed in short-lived species, such as OH and ozone, as a consequence of the January 2005 SPE. Figure 1 shows the MLS OH measurements (Figure 1a) together with WACCM (Figure 1b) and WACCM-D (Figure 1c) model predictions for 1424 January 2005 in the latitude band 6082.5N. Both MLS and the models show large OH enhancement during the SPE. The observed and modeled increase of OH occurred on 1718 January at altitudes between 60 and 82 km. In general, WACCM and WACCM-D agree well with observations. However, it is clear from Figure 1 that WACCM predictions overestimate OH values by about",{"entities":[[145,148,"INSTRUMENT"],[308,311,"INSTRUMENT"]]}]
["1. Introduction In the polar region, energetic particle precipitation (EPP) into the Earths atmosphere disturbs the chemical composition of the upper stratosphere, mesosphere, and lower thermosphere. Energetic (>10 MeV) solar protons are released from the Sun as a result of solar ares and coronal mass ejections and can gain direct access to Earths atmosphere in polar regions, guided by magnetic eld lines, which are basically open near the poles. In contrast, during periods of geomagnetic disturbances such as substorms and storms, low-energy electrons, with energies of 110 keV, precipitate from the magnetosphere, and high-energy electrons, with energies of 10 keV to a few MeV, precipitate from the inner magnetosphere and/or plasma sheet [Brasseur and Solomon, 2005; Horne et al., 2009; Lam et al., 2010].",{"entities":[]}]
["(425 K) and one above it (690 K for run UK, 673 K for run EC), are analyzed in detail between 1992 and 2001 (Figures 4 and 5), years when both model runs are available. The limited number of matches in 1996 for run EC results from a poor temporal correspondence of sampling and model results for that year. The higher frequency of comparisons with run UK result from the higher temporal resolution of the run UK output (every 24 hours). Run EC was run with output for every other day. It is important to note that Figures 4 and 5 have different vertical resolutions to accommodate the range of ozone mixing ratios at the two different isentropic levels.",{"entities":[]}]
["Jet Propulsion Laboratory JPL (2011), Chemical Kinetics and Photochemical Data for Use in Atmospheric Studies, vol. 15, JPL Publication, 06-2,",{"entities":[]}]
["EGU Journal Logos (RGB)Advances in GeosciencesOpen AccessNatural Hazards and Earth System SciencesOpen AccessAnnales GeophysicaeOpen AccessNonlinear Processes in GeophysicsOpen AccessAtmospheric Chemistryand PhysicsOpen AccessAtmospheric Chemistryand PhysicsOpen AccessDiscussionsAtmospheric MeasurementTechniquesOpen AccessAtmospheric MeasurementTechniquesOpen AccessDiscussionsBiogeosciencesOpen AccessOpen AccessBiogeosciencesDiscussionsClimate of the PastOpen AccessOpen AccessClimate of the PastDiscussionsEarth System DynamicsOpen AccessOpen AccessEarth System DynamicsDiscussionsGeoscientificInstrumentation Methods andData SystemsOpen AccessGeoscientificInstrumentation Methods andData SystemsOpen AccessDiscussionsGeoscientificModel DevelopmentOpen AccessOpen AccessGeoscientificModel DevelopmentDiscussionsHydrology and Earth SystemSciencesOpen AccessHydrology and Earth SystemSciencesOpen AccessDiscussionsOcean ScienceOpen AccessOpen AccessOcean ScienceDiscussionsSolid EarthOpen AccessOpen AccessSolid EarthDiscussionsThe CryosphereOpen AccessOpen AccessThe CryosphereDiscussionsNatural Hazards and Earth System SciencesOpen AccessDiscussions2312",{"entities":[]}]
["tigate the ozone budget. This marked the first time that direct measurements of chlorine species had been incor porated into a calculation of upper stratospheric 03. The calculations employed a newly developed iteration tech nique using the Mainz photochemical box model to derive the diurnal dependence of all important chemical trace gases, including species not measured by HALOE. The integration of chemical species over 1 day was repeated until the iteration converged to a diurnal cycle of chemical species consistent with the HALOE sunrise or sunset observations. ",{"entities":[]}]
["Figure 15. Comparison of MLS version 3 water vapour with the version 2 data from the ground-based mi crowave instrument. Data are from the period of Jan uary 23, 1992 to October 13, 1992; a total of 186 days for which both MLS and ground-based measurements were available. (left) Mean profiles; the solid curve is MLS data, the dashed curve ground-based. The (right) Mean difference, MLS ground based (solid) and the rII1S cunerence ",{"entities":[[27,30,"INSTRUMENT"],[240,243,"INSTRUMENT"],[335,338,"INSTRUMENT"],[406,409,"INSTRUMENT"]]}]
["The assimilation runs with and without GFAS emissions do not show a signicant difference for D+0 hindcasts, illustrating the dominating impact of the initialization by assimilation of IASI observations.",{"entities":[[184,188,"INSTRUMENT"]]}]
["Figures 9 and 10 show the comparison between SCIAMACHY and MLS at midlatitudes for the period August 2004April 2012. At northern midlatitudes, the MLS O3 trend prole has less variability with altitude than that of SCIAMACHY. There are some signicant deviations between 30 and 35 km, with SCIAMACHY showing negative trends while MLS shows small positive trends. Positive trends seen by SCIAMACHY around 40 to 45 km agree with",{"entities":[[45,54,"INSTRUMENT"],[59,62,"INSTRUMENT"],[147,150,"INSTRUMENT"],[214,223,"INSTRUMENT"],[288,297,"INSTRUMENT"],[328,331,"INSTRUMENT"],[385,394,"INSTRUMENT"]]}]
["Lutman, E. R., R. Toumi, R. L. Jones, D. J. Lary, and J. A. Pyle, Box model studies of C1Ox deactivation and ozone loss during the 1991192 Northern Hemisphere win ter, Geophys. Res. Lett., 1415-1418, 1994. ",{"entities":[]}]
["Pickett, H. M., W. G. Read, K. K. Lee, and Y. L. Yung (2006b), Observation of night OH in the mesosphere, Geophys. Res. Lett., 33, L19808, doi:10.1029/2006GL026910.",{"entities":[]}]
["Averages Jan2004-Mar2004-1001020304050SO2 [pptv]1214161820Altitude [km]ACE-FTSMIPASBias-100-50050100Diff. [pptv]1214161820Scatter020406080100Diff. [pptv]1214161820rms(bias)comb. precis.No. of pairs0501001502001214161820Averages Jul2005-Sep2010 excl.volc.-1001020304050SO2 [pptv]1214161820Altitude [km]ACE-FTSMIPASBias-100-50050100Diff. [pptv]1214161820Scatter020406080100Diff. [pptv]1214161820rms(bias)comb. precis.No. of pairs01000200030001214161820M. Hpfner et al.: MIPAS SO2 in the UTLS",{"entities":[]}]
["Imager (MVIRI), Geostationary Operational Environmental Satellite (GOES) and Spinning Enhanced Visible and InfraRed Imager (SEVIRI) instruments. However, it is probably fair to say that these have been conspicuously underused in global NWP compared to the use of UV backscatter data. There are a number of reasons for this. Firstly, the ozone information comes from IR channels that have a signicant sensitivity to clouds and surface emission. Thus errors in the detection of cloud contamination and the characterization for surface emissivity and skin temperature can potentially undermine the useful ozone signal in the data. Secondly, the observed radiances and the radiative transfer (RT) model used in the interpretation of the observations are both prone to biases. This problem is of course not unique to ozone-sensitive data and applies to all radiance observations. However, there is signicantly less unbiased independent ozone information available for the diagnosis and correction of systematic errors in ozone-sensitive radiances compared to that available to detect and correct biases in temperature-sensitive radiances.",{"entities":[[8,13,"INSTRUMENT"],[16,65,"SPACECRAFT"],[67,71,"SPACECRAFT"],[77,122,"INSTRUMENT"],[124,130,"INSTRUMENT"]]}]
["For 10 years, the development of these monitoring and forecasting abilities has been the primary goal of a series of European projects. The European Union project MACCII (Monitoring Atmospheric Composition and Climate Interim Implementation) was the third in a series of projects funded since 2005 to build up the atmospheric service component of the Global Monitoring for Environment and Security (GMES)/Copernicus European programme (Peuch et al., 2014). In this paper, the term MACC refers to both the MACC and MACC-II projects. The nal goal of MACC is to cover all aspects of atmospheric dynamics and chemistry with one global data assimilation system (DAS) based on an operational numerical weather prediction (NWP) system.",{"entities":[]}]
["Table 2 Magnetically perturbed O2 line frequencies (in MHz) and strength relative to zero =eld values",{"entities":[]}]
["Do rnbrack, A., Birner, T., Fix, A., Fientje, H., Meister, A., Schmid, H., Browell, E.V., Mahoney, M.J. Evidence for inertia gravity waves forming polar stratospheric clouds over Scandinavia. J. Geophys. Res. 107 (20), 8287, doi:10.1029/2001JD000452, 2002.",{"entities":[]}]
["[20] During 20 January (the SSW onset period; Figure 2b), the polar vortex has become elongated and varies considerably with altitude. The predominantly westerly (i.e., circumpolar) flow seen during 10 January has now evolved into a flow with considerable meridional (northsouth) excursions. The zonal mean wind is weakened at nearly all altitudes, as suggested in Figure 1a. In the simulations, GWs",{"entities":[]}]
["Rodger, C. J., Carson, B. R., Cummer, S. A., Gamble, R. J., Clilverd, M. A., Sauvaud, J.-A., Parrot, M., Green, J. C., and Berthelier, J.-J.: Contrasting the efciency of radiation belt",{"entities":[]}]
["Wheeler, M., and G. N. Kiladis (1999), Convectively coupled equatorial waves: Analysis of clouds and temperature in the wavenumber-frequency domain, J. Atmos. Sci., 56, 374 399.",{"entities":[]}]
["600 MHz tunable between approximately 230 and 280 GHz. By means of the observed line shape, together with pressure and temperature vertical profiles, a mathematical deconvolution process exploits the pressure broadening of spectral lines to determine the emitting molecules concentration as a function of altitude from about 15 to 80 km. For water vapor, only the integrated column density can be obtained [Fiorucci et al., 2008].",{"entities":[]}]
["As noted in Sect. 2, there is a discrepancy between simulated and observed SWOOSH tropical lower-stratospheric ozone post-2005. This is also a region where there is considerable divergence between different observational estimates of ozone changes (WMO, 2014). Both of these factors (modeldata differences and observational uncertainty) motivated the additional analysis with exclusion of ozone changes in the tropical lower stratosphere from our S/N analysis. In the tropics excluded case, there is high condence in our detection of the model ODS signal in the lower stratosphere. For the two reasons outlined above, we have less condence in the interpretation of our S/N results for the global tropics included case, especially for method 2 for which the PC time series are used. Due to the noticeable divergence between simulated and observed post-2005 ozone changes in the tropical lower stratosphere, inclusion of the tropics re-",{"entities":[]}]
["[8] In the following, we first discuss in section 2 the situation of the upper troposphere/lower stratosphere from the perspective of water vapor. Section 3 describes data and methods, and section 4 compares stratospheric water vapor concentrations from model predictions with observations. Section 5 presents an analysis of errors arising from the method to evaluate the AC paradigm, allowing a quantification of the effect of physical processes deliberately neglected in the AC paradigm. Finally, section 6 provides a discussion of the implications of our results.",{"entities":[]}]
["upper stratosphere. The easily photolyzed sources also have a shorter lifetime and respond faster to emission changes at the surface. For example, CH3CC13 has begun to decline, while CF2C12 has not. While there are some variations in the data that are not readily explained, the model has some limitations which do not allow for optimum simulation of year-to-year variations or coupling of dynamics, radiation, and chemistry. ",{"entities":[]}]
["Two examples are shown to illustrate the performance of the PSC detection algorithm. The top panel of Fig. 7 shows the CALIOP 532-nm backscatter coefcient data from a single orbit on 15 June 2006. The corresponding PSC mask produced from the PSC detection algorithm is shown in the bottom panel of Fig. 7. The averaging scale (5-, 25-, or",{"entities":[[119,125,"INSTRUMENT"]]}]
["grant 1-1252. We thank the European Center for Medium-Range Weather Forecasts for their provision of grid point data for both the PEMWA and the PEMWB missions. ",{"entities":[]}]