1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 Analysis of Air Transport and Oxidation Chemistry in the NorthAtlantic Region from Interpretations of Non-Methane Hydrocarbon (NMHC) Measurements at Pico Mountain, Azores D. Helmig1, D. Tanner1, R.C. Owen2 and R. E. Honrath2 and D. Parrish3 1 Institute of Arctic and Alpine Research (INSTAAR), University of Colorado, Boulder, CO 80309, USA 2 Department of Civil and Environmental Engineering, Michigan Technological University, Houghton, Michigan, USA 3 Chemical Sciences Division, National Oceanic and Atmospheric Administration, Boulder, CO 80303, USA Manuscript in preparation for Journal of Geographical Research June 27, 2006 Abstract 20 21 One year of continuous measurements of non-methane hydrocarbons at the mountaintop PICO- 22 NARE observatory on Pico Island, Azores were used to investigate seasonal oxidation chemistry 23 and transport patterns in the central North Atlantic Region. NMHC at this site exhibited seasonal 24 and short-term variations and cycles that reflect the distance of the island from continental 25 sources of NHMC emissions and oxidation of NMHC by the seasonally highly variable OH 26 radical. Substantially enhanced NMHC levels during the summer of 2004 were attributed to the 27 impact of long-range transport of biomass burning plumes resulting from Northern Canada and 28 Alaskan wildfires. 29 NMHC concentrations and their relative ratios were valuable in identifying transport situations 30 where anthropogenically influenced air from the mid and western U.S. was transported to Pico in 31 5-8 days. 32 processing (‘photochemical clock’) was shown to yield results that were in qualitative agreement 33 with trajectory and FLEXPART analysis interpretations. 34 measured at PICO-NARE throughout all seasons. Enhanced ozone levels were observed in air 35 that had relatively ‘fresh’ photochemical signatures (e.g. ln [propane]/[ethane] > -2.5). Ozone at During summer, air samples were the most photochemically processed. Interpretations of NMHC ratios for use as a relative scale for photochemical Ozone in excess of 35 ppbv was 1 36 lower levels (< 40 ppbv) was always correlated with more processed air patterns (‘older’ air with 37 ln [propane]/[ethane] < -2.5. 38 39 1. Introduction 40 41 Non-methane hydrocarbons (NMHC) in the atmosphere show considerable variations on spatial 42 and temporal scales, their concentrations being determined by the strength of emission sources 43 and atmospheric removal processes. Atmospheric oxidation is mostly due to reaction with the 44 OH radical, with reaction rate constants increasing significantly with the molecule size. 45 Consequently, lighter hydrocarbons (having the slowest reaction rates) exhibit much longer 46 lifetimes and their atmospheric concentrations decline at slow enough rates for NMHC 47 concentrations to remain high enough after several days of transport to impact air chemistry at 48 remote downwind locations. Since many individual NMHC have common emission sources and 49 since their emission ratios vary comparatively little, changes in absolute concentrations and 50 NMHC ratios can be used as tools to decipher atmospheric transport and oxidation chemistry. 51 Several researchers have investigated this utility and have presented a framework for the 52 interpretation of atmospheric light (C2-C6) hydrocarbon observations. 53 54 Selected NMHC can be used as tracer for specific emission sources or events. For instance, 55 isoprene is a selective tracer for biogenic emissions (Fehsenfeld et al., 1992), acetylene has been 56 found to be significantly enhanced in biomass burning plumes (DeGouw et al., 2004) and light, 57 saturated and unsaturated NMHC (e.g. ethane, propane) have been used to identify influences 58 from urban energy use and petrochemical industries (Blake and Rowland, 1995; Jobson et al., 59 2004). Diurnal concentration changes of light, unsaturated NMHC (ethene, propene) allowed 60 identifying occurrences of upslope and downslope flow conditions at Mauna Loa Observatory 61 (Greenberg et al., 1996). 62 63 Certain NMHC (e.g. butanes) have similar atmospheric removal rates and hence their, 64 atmospheric ratios show little variations during atmospheric transport and processing. Their 65 analytical data can therefore be used as a quality control tool in NMHC measurements (Parrish et 66 al., 1998). 2 67 68 The variability of NMHC concentrations can provide information on the impact or distance of a 69 measurement site from pollution sources. A ‘remoteness’ scale has been proposed, that is 70 derived from a plot of the natural logarithm (ln) of the standard deviation of ambient NMHC 71 concentrations at a given site versus their estimated lifetime (Jobson et al., 1999). Relative 72 changes of the ratio of branched versus straight n-alkanes have been used to infer the importance 73 of halogen and nitrate radical versus OH radical chemistry as the reaction rates of these two 74 different oxidation routes are significantly different enough to cause shifts in the atmospheric 75 concentration ratios of these isomeric compounds (Penkett et al., 1993; Finlayson-Pitts, 1993). 76 77 NMHC ratios and concurrent measurements of ozone were also applied to investigate potential 78 changes in the oxidation chemistry of the atmosphere. In particular, the relative increase of 79 ozone at observed NMHC ratios was used as an argument for an increased ozone production 80 (respectively reduced ozone loss rates) in long-range transport across the Pacific Ocean (Parrish 81 et al., 2004). NMHC measurements from remote, marine environments have also been applied 82 for estimating mean OH radical fields during transport of air in the marine boundary layer and 83 lower free troposphere (Ehhalt et al., 1998; Williams et al., 2000). 84 85 The possibilities for using NMHC data for interpretations of atmospheric oxidation processes are 86 particularly promising in situations where observations can be obtained in air that has traveled 87 for extended periods of time without being influenced by recent emissions or surface processes. 88 Hence, remote islands that are high enough to probe free tropospheric air are ideal locations for 89 this research. These aforementioned considerations motivated the monitoring of NMHC at the 90 mountaintop PICO-NARE site on Pico Island, Azores. These measurements commenced in the 91 summer of 2004 and have been continuous since during most times when the station was on 92 power. In this paper we present data from the first year of these new observations and examples 93 of interpretations that demonstrate the utility of the NMHC data for interpretations of oxidation 94 and transport processes in the North Atlantic region. 95 96 2. Methods 97 3 98 2.1. PICO-NARE Station 99 100 The PICO-NARE observatory is located in the summit caldera of the inactive Pico Mountain 101 volcano (38.47N, 28.40W), the highest mountain on Pico Island, and in the Azores, Portugal. At 102 2225 m asl, lower, free tropospheric air is sampled at the station during most times. More 103 information on the geo- and topographical features are provided by Kleissl et al. (2006). A 104 detailed analysis of boundary layer height, and of mechanical uplifted and buoyant flow 105 conditions showed that air from lower elevations was potentially lifted up to the station height up 106 to 50% of the days during some months, but to a lesser extent (~ 25% of the days) during the 107 summer. However, chemical measurements of nitrogen oxides and carbon monoxide at the 108 observatory showed very little (resp. negligible) influence from island emission sources even 109 during upslope conditions, which infers that even during uplifting events, mostly lower 110 tropospheric air is transported to the station (Kleissl et al., 2006). Data and interpretations from 111 other research, including studies of oxidized nitrogen species, ozone, carbon monoxide and of 112 aerosol properties at PICO-NARE have been presented previously (Honrath et al., 2004; Fialho 113 et al., 2006) and in other contributions to this special issue (Val Martin et al., 2006). 114 115 2.2 NMHC Measurements 116 117 The remoteness of the PICO-NARE site and the limitations for power and for supply of cryogen 118 and consumable gases determined the design of an analytical system that was tailored towards 119 this unique situation. All consumable gases and blank air were prepared at the site with low- 120 power gas generators. The instrument was designed to follow automated startup and shuddown 121 procedures and could be remotely controlled from our Boulder, CO offices. Ozone was removed 122 by flowing the sample air through an ozone scrubber prepared from sodium-thiosulfate- 123 impregnated glass wool. After sample drying and NMHC focusing on a mulit-stage solid 124 adsorbent trap, NMHC were analyzed by thermal desorption with gas chromatography (GC) 125 separation and flame ionization detection (FID). The instrument was calibrated by regular 126 injections of a multi-component, rural air standard that was quantified prior to shipment against a 127 gravimetric hydrocarbon standard scale in the NOAA Aeronomy Laboratory. A second, remote, 4 128 ambient air standard (collected at Niwot Ridge, Colorado) was injected every 3-4 days for 129 quality control. 130 131 Sample volumes of 600 ml (10 min collection time) and 3000 ml (50 min collection time) were 132 alternated for quantification of ethane and NMHC > C2, respectively. Typically, a total of 12 133 ambient air samples, one standard and one blank sample were analyzed daily. Data were 134 transferred daily for instant quality control and analysis. The primary calibration standard was 135 returned to Boulder in spring 2006 and the control analysis on the independently calibrated 136 NOAA GC system showed that C2-C6 NMHC mixing ratios were within +/- xx % (DAVID, WE 137 PROBABLY WILL HAVE TO PUT OUR BEST GUESS NUMBERS HERE AS WE STILL 138 DON’T HAVE THIS ANALYSIS BACK FROM PAUL? ANY UPDATE ON THAT? WE 139 DEFININTELY NEED TO HAVE THIS DONE BY THE TIME THE REVIEWS COME 140 BACK) of the values determined two years earlier, prior to the shipment to Pico. NMHC were 141 quantified using compound-specific FID response factors, as determined from the primary 142 standard injections. Quantified NMHC included ethane, propane, n-butane, i-butane, i-pentane, 143 n-pentane, n-hexane and isoprene (the ethane record doesn’t begin until in fall 2004 when some 144 modifications in the focusing procedure allowed its quantitative analysis). 145 described experiments analytical precision and accuracy were estimated to be better than 10% 146 for mixing ratios > 100 ppt and approximately a factor of 2 higher for levels between the 147 detection limit (which typically were ~ 30, 11, and 1-2 148 respectively) and 100 pptv. More instrumental details have been provided elsewhere (Tanner et 149 al., 2006). From the above pptv for C2, C3, and C4-C6, 150 151 2.3 Trajectory Analysis 152 153 Backward trajectories were calculated with the Hybrid Single-Particle Lagrangian Integrated 154 Trajectories (HYSPLIT) model (Draxler and Rolph, 2003). HYSPLIT uses meteorological data 155 from the National Weather Service’s National Center for Environmental Prediction (NCEP) final 156 analysis (FNL). Data are available at 6-hour resolution with 13 pressure altitude levels. A set of 157 six trajectories were calculated, one terminating at the station, four terminating at grid points 5 158 adjacent to the station and separated from it by 1o, and one terminating directly below the station 159 at 2000 m asl.. Trajectories were run 10 days backward in time. 160 161 2.4 FLEXPART Simulations 162 163 Besides the trajectory analysis, the FLEXPART particle dispersion model (versions 5.2 and 6.2, 164 (Stohl et al., 1998; Stohl and Thomson, 1999; Stohl et al., 2005)) was used to evaluate derived 165 NMHC transport times with synoptic transport modeling results. FLEXPART version 6.2 was 166 driven with data from the European Centre for Medium Range Weather Forecasts (ECMWF) 167 (ECMWF, 2005) with a 1 degree horizontal resolution, 60 vertical levels and a temporal 168 resolution of 3 hours, using meteorological analyses at 0000, 0600, 1200, and 1800 UTC, and 169 ECMWF 3-hr forecasts at intermediate times (3, 9, 15, 21 UTC). FLEXPART version 5.2 was 170 driven with data from wind fields from the NOAA NCEP FNL. The FNL data were downloaded 171 from the National Center for Atmospheric Research data archive, available every 6-hours with a 172 horizontal grid spacing of 1 1 , and 21 vertical levels between 1000 and 100 hPa 173 174 Version 5.2 of the model was run in its forward mode to simulate CO enhancements at the PICO- 175 NARE station resulting from the transport of North American and Asian emissions. These 176 emissions were divided into one day age classes, which was also useful for determining the time 177 since emission. CO emissions were released into the lowest 300 m of the atmosphere over North 178 America and Asia. Emissions were based on the EDGAR 3.2 Fast Track 2000 dataset [Olivier et 179 al., 2001] for anthropogenic sources only with a 1 degree resolution. 180 181 Version 6.2 of the model was run in its backward mode to create “retroplumes”, similar to 182 backward trajectories. Retroplumes are simulated from the release of thousands of particles at 183 the receptor that are advected backwards in time. Retroplumes are superior to trajectories in that 184 they allow for an assessment of the deformation of an air mass as it travels and for determining 185 source regions for observed enhancements (Seibert and Frank, 2004). 186 initiated every three hours with 20,000 particles released over a three hour time interval into a 1 187 degree x 1 degree grid box centered on the PICO-NARE station, over an altitude range of 1750 188 m asl to 2750 m asl. Particles were followed backward in time for 20 days. Retroplumes were 6 189 190 3. Results and Discussion 191 192 3.1 NMHC Mixing Ratios 193 194 Plots with the individual sample data (representing a total of 1958 analyzed air samples) for 195 ethane, propane, and n-butane from Aug. 2004–Sept. 2005 were presented by Tanner et al. 196 (2006). Here we combined these data to monthly whisker plots that show the minimum, 5, 25, 197 50, 75, and 95 percentile, and the maximum values of measured mixing ratios during each month 198 of available measurements (Fig. 1). Sinusoidal best fit curves were calculated from the (diurnal 199 resolution) data and are included in these graphs to illustrate a smoothed seasonal NMHC cycle. 200 201 These data show the typical Northern Hemisphere seasonal cycle of NMHC with lower mixing 202 ratios in the summer and maximum values in late winter. This behavior to a large extent is 203 driven by the annual concentration changes of the OH radical, which is closely linked to the 204 latitudinal solar radiation cycle. High variability in NMHC mixing ratios was observed at any 205 given time of year. It is noteworthy that all of these features show relations and dependencies 206 toward the individual NMHC reactivity with OH and the resulting NMHC lifetime. The longest- 207 lived NMHC, ethane, shows the relatively smallest amplitude between the mean winter and 208 summer mixing ratio and the smallest relative variability on short (e.g. weeks) time scales. All 209 of these features increase with increasing molecule size (respectively shorter OH lifetime). The 210 seasonal maximum and minimum of ethane occurs the latest of all compounds (March 3 and 211 September 3, respectively), as due to its slower OH reaction, ambient levels respond with a 212 longer delay to the seasonal OH cycle. Heavier NMHC were found to maximize earlier, up to 213 around January 20 for the most reactive compounds, and also had their seasonal minimum 214 earlier, around July 18. These features in the Pico NMHC data are in agreement with data from a 215 number of other sites, which, along with their seasonal OH dependencies, have been presented 216 and discussed in detail in the literature (e.g. Jobson et al., 1994; Goldstein et al., 1995; Gautrois 217 et al., 2003). 218 7 219 Comparison of the NMHC >C3 for summer 2004 with data from the corresponding period 220 during 2005 shows a higher variability as well overall higher mixing ratios during 2004. As 221 discussed in detail in other contributions to the ICARTT issue, the summer of 2004 was 222 characterized by an unusually high occurrence of boreal wild fires in Northern Canada and 223 Alaska, outflow of which was frequently observed at Pico (Val Martin et al., 2006). Substantial 224 enhancements in NMHC, at times increasing to twice their seasonal background levels, were 225 observed in these boreal fire plumes. Overall, during times when the station was impacted by 226 boreal fire plumes (as defined in Val Martin et al., 2006), e.g. propane 25/50/75 percentile 227 mixing ratios were 65/81/143 pptv, whereas outside of fire events, they were 25/51/87 pptv 228 during the 2004 summer. 229 230 A number of data sets have been presented in the literature that allow comparisons with the 231 PICO-NARE measurements. We included two related data series in Figure 1, notably from two 232 years of measurements at the continental, remote boreal site in Fraserdale, Ontario (50oN, 82oW) 233 (Jobson et al., 1994) and from the Mauna Loa Observatory Photochemical Experiment-2 (19oN, 234 xxoW) (Greenberg et al., 1996), which, as shown below, bracket the Pico measurements and 235 allow a further interpretation of the particular conditions encountered at Pico. 236 237 Probably the most related study are the MLOPEX-2 measurements. Similar to PICO-NARE, the 238 Mauna Loa site is a remote mountaintop island location, where, during downslope conditions, 239 free tropospheric air is sampled that has traveled over the ocean for several days. 240 more prominent diurnal upslope-downslope cycle and data presented by Greenberg et al. were 241 broken up into the occurrences of these two flow regimes. 242 25/50/75 percentiles for the time periods spanned by the width of the boxes. The shown MLO 243 data are from downslope (e.g. free tropospheric air) conditions. Upslope data for MLO typically 244 were higher, with relative enhancements increasing with decreasing molecule liftetime. 245 MLOPEX-2 data are consistently lower for all NMHC and during all seasons. The differences in 246 Pico and MLO NMHC mixing ratios increases with decreasing lifetime, e.g. while ethane mixing 247 ratios compare to within ~20%, n-butane values at MLO are more than 5 times lower than at 248 Pico. 249 experiences overall higher NMHC values than Pico. Again, differences in these two data series MLO has a Included data in Figure 1 are the In contrast to MLO, Fraserdale, a remote, low elevation continental forest site, 8 250 become more pronounced with molecular weight, although this time the PICO-NARE data are 251 the ones becoming increasingly lower. 252 253 Higher NMHC levels at Fraserdale than at Pico, and higher NMHC levels at PICO-NARE than 254 at MLO are likely due to several reasons. 1. Aircraft profiles have shown that NMHC mixing 255 ratios generally decline with height within the free troposphere (e.g. Blake et al., 1997), with 256 larger concentration changes being observed for shorter-lived compounds. Pico is higher than 257 Fraserdale, and MLO is at about 1200 m higher altitude than the PICO-NARE station, 258 consequently NMHC mixing ratios would be expected to be highest at Fraserdale, followed by 259 PICO-NARE and MLO. Secondly NMHC mixing ratios in the lower troposphere decrease 260 towards lower latitude (Rudolph, 1995). Again, this dependency would infer higher NMHC 261 levels at Fraserdale (50oN), followed by Pico (38oN) and MLO (19oN). Of further importance is 262 the distance to the adjacent continents, which is about two times as much for MLO, and which 263 will cause transport and photochemical processing times from continental sources to be longer, 264 resulting in more depleted NMHC ratios at MLO compared to PICO-NARE. 265 comparisons of the PICO-NARE data with other data sets from higher (than Pico) northern 266 latitudes in Canada, the Atlantic Region and Europe (as summarized by Gautrois et al. (2003)) 267 shows that Pico NMHC levels are unanimously lower, both during the winter and in the summer 268 compared to these locations. Further 269 270 The cumulative distribution of NMHC during fall 2004 (September 22 to December 20), winter 271 2004-2005 (December 21 to March 19), spring 2005 (March 20 to June 20) and summer 2005 272 (June 21 to September 21) is shown in Figure 2. Data series that do not extend in the lower 273 percentage range resulted from respective fractions of these data being reported below the 274 instrument detection limit (for instance, i- and n-pentane were below the detection limit in about 275 4% of the measurements during fall 2004, whereas during summer 2005, ~85% and 60% of 276 chromatograms had i- and n-pentane peaks that were too small to quantify). The regression line 277 slopes through these individual data series indicate the variability of the atmospheric 278 concentrations of a given compound. Steeper slopes are observed for longer-lived NMHCs (e.g. 279 ethane) as these compounds have a higher atmospheric background concentration which reduces 280 the relative variability caused by emission influences. It is noteworthy that regression line slopes 9 281 are lower for the summer, which likely can be attributed to the shorter seasonal atmospheric 282 lifetime and resulting lower background concentrations, which will cause higher relative 283 variabilities. 284 285 Results for isoprene measured at the station were presented by Kleissl et al. (2006). Isoprene, 286 was typically not detected (< 1 pptv) in winter and nighttime samples. During spring, isoprene 287 was occasionally observed in samples collected during morning to evening hours. Occurrences 288 and mixing ratios of isoprene increased towards late summer. 289 was detected on 60% of all days in the afternoon with maximum mixing ratios reaching up to 27 290 pptv. The isoprene data clearly show seasonal and diurnal dependencies that are determined by 291 both the expected seasonal changes in isoprene emission rates from vegetation growing at lower 292 elevation on Pico Island and by occurrences of buoyant and mechanical uplift flow that 293 transports air from lower parts of Pico to the observatory (Kleissl et al., 2006). During August 2005, isoprene 294 295 3.2 NMHC Variability 296 297 The variability of NMHC during each of the four differentiated seasons is directly related to the 298 slopes of regression lines through the data in Figure 2. On a cumulative distribution plot, the 299 slope of the regression line is the standard deviation of the data, assuming the data (including 300 points falling below the detection limit) are log-normally distributed. The seasonal lifetimes of 301 NMHC were determined from their OH reaction constants and seasonal OH radical 302 concentration. [OH] was estimated at a 1-day resolution according to (Goldstein et al., 1995): 303 304 [OH] = A [1-B cos(2*pi * t/365)], 305 306 where A=1.6*106 and B=0.80. 307 (Spivakovsky et al., 2000) for 800 hPa, 36.0oN, 27.5oW. Daily OH concentrations were then 308 averaged to seasonal OH values within the defined time periods. Reaction rate constants were 309 adjusted to the temperatures measured at the PICO-NARE station during the respective season. 310 Please note that this local lifetime represents an estimate for the conditions at the receptor site; 311 the actually encountered lifetime during transport of a NMHC to Pico may have deviated from This calculation utilizes monthly average OH values 10 312 this estimate dependant on the geographical and atmospheric conditions during the transport 313 path. Also note that this analysis in not very sensitive towards the applied [OH], but more so 314 towards the relative reactivity differences between individual compounds. Consequently, errors 315 in the estimated, total [OH] will only have little effect on the results for the regression 316 coefficients (Jobson et al., 1999). 317 and for each of the four seasons. The results shown in Figure 3 show well correlated linear 318 relationships for all four seasons. Regression line slopes, according to lnx = A -b. for each 319 seasonal data set are included in the figure. Seasonal differences in the b-values were not 320 statistically significantly different at P > 95% for seasons where data of all C2-C5 NMHC were 321 included. The best fit linear regression analysis through the data for all seasons yielded lnx = 322 1.60 -0.44. lnx – lifetime estimates were obtained for each compound 323 324 The exponent b in this equation has been noted to describe the importance of sink terms in the 325 regional variability budget whereas the coefficient A can be related to the degree of 326 photochemical aging; A-values have been used to derive estimates of transit times for different 327 sample sets (Jobson et al., 1990). 328 329 Interpretation of observed values for b from different sites has shown that b approaches 0 near 330 urban areas, where the variability is strongly influenced by differences in the strength of local 331 emission sources, whereas b-values close to 1 are found in stratospheric data sets, where the 332 variability is low and dominated by chemical loss alone. The mean Pico value of 0.44 +/- 0.03. 333 compares well with data from three aircraft data sets collected over other diverse remote areas, 334 including the Arctic Boundary Layer (ABLE3A), the equatorial Atlantic (TRACE-A) and the 335 western Pacific (PEM-West B) experiment, which resulted in b-values of 0.46–0.53 (ref). This 336 comparison illustrates a rather high similarity between the continuous, seasonal Pico data and the 337 results from the comparatively short aircraft campaigns that have been previously presented in 338 the literature. 339 340 3.3 Ratios of NMHC 341 11 342 Correlation plots of all C2-C6 NMHC as well as of these NMHC with CO are shown in Figure 4 343 and results for the linear regression analyses are given in Table 1. The common feature in these 344 data is that regression line slopes of (NMHCA/NMHCB with carbon number NMHCA < NMHCB 345 increase monotonically with increasing carbon number of NMHCB. This behavior is expected as 346 the atmospheric mixing ratios (and lifetimes) of NMHC generally decrease with increasing 347 carbon number. For an individual pair of NMHC, the regression line slopes become larger 348 towards the summer, as the longer-chain NMHC are removed from the atmosphere faster than 349 the more stable, shorter-chain NMHC. Regression coefficients generally decrease towards the 350 summer, as shorter liftetimes, lower concentrations and higher relative variability cause the 351 correlation between individual compounds to become weaker. In general, compounds with 352 similar lifetimes generally show better correlations than compound pairs with much different 353 lifetimes. However, the correlation between CO and ethane (which have very similar lifetimes) 354 is notably weaker, likely because of their differences in primary and secondary sources. 355 Although fossil fuel combustion is a common source of both gases, CO is also a degradation 356 product of hydrocarbons (mainly methane) in the atmosphere. 357 358 The OH reaction rate constants of the isomeric pairs iso-butane and n-butane, and of iso-pentane 359 are very similar and consequently, the atmospheric ratios of these two compound pairs is 360 expected to change very little during transport and photochemical oxidation. Their correlations 361 in the data from Pico, differentiated by the four seasons, as well as their ratio against the absolute 362 levels of n-butane and iso-pentane, respectively, are shown in Figure 5. The tight correlation 363 between these two compounds is very obvious. Deviations and larger scatter at lower (e.g < 50 364 pptv) mixing ratios to some extent can be attributed to the loss of precision when mixing ratios 365 approach the detection limit. 366 367 For the butanes, no statistically significant difference was found in the regression line slope 368 between the four seasonal data series. The regression line slope for all data was calculated to be 369 0.51 +/- 0.01 (R2 = x.xx). Similar values (range 0.37–0.55) have been reported in data from a 370 multitude of other sites in both continental and marine environments (e.g. Bottenheim and 371 Shepherd, 1995; Bottenheim et al., 1997; Greenberg et al., 1996; Parrish et al., 1998). The same 372 analysis yielded a slope of slope of 0.69 +/- 0.08 (R2 = 0.93) for the n-pentane/i-pentane data. 12 373 The graphs on the right side of Figure 5 investigate possible changes in the oxidation chemistry 374 of these compound pairs by season as well as by their absolute concentrations. Other than for the 375 increase in scatter at lower mixing ratios, the butanes do not show any systematic seasonal 376 changes. The pentane plot looks somewhat different, as higher n-pentane/iso-pentane ratios are 377 observed for the spring and summer data. This point is also visible in the cumulative distribution 378 plots (Fig. 2) where, only for the summer data, the i-pentane data distribution falls above the n- 379 pentane values. This behavior points to either different source region emission ratios or to 380 different oxidation chemistry during the summer months. Changes in n-pentane ratios have 381 previously been investigated by several other researchers. 382 pentane ratios during summer months and in low-concentration (well aged) samples were also 383 evident in the Fraserdale data (Jobson et al., 1994) as well as during ICARTT and during 384 NEAQS from the Ron Brown (Parrish xxxx). Notably, the n/i-pentane ratios in the Pico data 385 overall seem to be higher than in the data from (respectively closer to) the continental regions. 386 This behavior is contrary to expected OH kinetics, as a relative decrease of n-pentane/i-pentane 387 would be expected during summer, due to the slightly higher n-pentane OH reaction rate 388 constant (3.94 x 10-12 cm3 molecule-1s-1 at 298 K) (Atkinson, 1994) compared to 3.9 x 10-12 cm3 389 molecule-1s-1 at 298 K ???? for iso-pentane (2-methylbutane) (Atkinson, 1989), which should 390 result to lower n/i-pentane ratios in summer. These dependencies may point towards seasonally 391 changing competition between alternative destruction pathways, such as by the NO3 radical or by 392 chlorine chemistry (Penkett et al., 1993). Similarly enhanced n-pentane/i- 393 394 As a note of caution, it should be pointed out that our current interpretations of the summer 395 pentane data are somewhat limited. Even though the cumulative distribution plot shows n- 396 pentane data about twice as high as i-pentane, the pentane ratio plot shows the summertime value 397 to be around 1, with quite a substantial degree of scatter. For i-pentane, only 14% of the summer 398 data, and for n-pentane, only 41% of the summer data were above the detection limit. Only 11% 399 of the samples had both i- and n-pentane. So, any conclusions about these data are from the 400 highest 11% of concentrations observed during the summer. Also, it may be possible that at such 401 low concentrations (< 5 ppt) increasing sampling or measurements artifacts have to be taken into 402 account. The pentane observations and their preliminary interpretations presented here are 13 403 nonetheless a motivation for future, more thorough and accurate studies of pentane chemistry at 404 Pico. 405 406 The distribution of NMHC data in a double natural logarithm plot of [n-buante]/[ethane] versus 407 [propane]/[ethane] (Figure 6) can be used to investigate the degree of photochemical processing 408 that occurred in air reaching Pico. Data in these plots is distributed between two theoretical lines 409 that are determined by the assumptions that air with a common ratio of these compounds at a 410 source would have only been altered by OH photochemistry (kinetic line) or by dilution with air 411 that has zero concentrations of both compounds of consideration (dilution line). The seasonal 412 differences in NMHC oxidation are clearly visible in these data. During winter, most data have 413 larger ratios and are less variable, indicative of less photochemical processing, and/or more 414 homogenously distributed source regions and air transport. In contrast, spring and summer data 415 are more scattered (weaker R2 values); the lower [NMHC]/[ethane] ratios are indicative of the 416 higher degree of air processing that occurred during transport. The regression through all data 417 yields a slope of 1.60 (±0.04), which is within the range of slopes reported for this analysis from 418 several other experiments (Parrish et al., 2004). DAVID P., ANYTHING ELSE YOU WOULD 419 LIKE TO ADD? 420 421 3.4 NMHC Processing and Ozone 422 423 In general understanding the temporal variability of tropospheric ozone at any particular location 424 is complex because several processes can have significant impacts, and these impacts vary 425 strongly on different time scales. In situ photochemical production and destruction proceed at 426 rates that vary with the ambient levels of ozone precursors and variables such as sunlight and 427 water vapor levels. Surface deposition and destruction by reaction with local emissions of NO or 428 reactive NMHC can drastically reduce near-surface ozone concentrations at rates that vary with 429 the characteristics of the planetary boundary layer and the flux of local emissions. Transport of 430 ozone to the site from the stratosphere or from upwind regions of strong photochemical 431 production can greatly increase ozone concentrations. 432 433 The PICO-NARE site is ideally situated to isolate the effects of the regional photochemical 14 434 production and destruction in the central North Atlantic from the effects of the other processes. 435 Kleissl et al. (2006) show that air sampled at the site is characteristic of the lower free 436 troposphere with essentially no opportunity for significant effects from surface deposition, 437 destruction by reaction with local emissions, or photochemical production from locally emitted 438 precursors. The varying influence of the transport of stratospheric ozone often dominates the 439 variability of ozone in free tropospheric data sets. Since ozone from the stratosphere has a steep, 440 negative correlation with CO (see e.g., (Danielsen et al., 1987)) the influence of stratospheric 441 ozone transport can be evaluated from the correlation of ozone with CO. Honrath et al. (2004) 442 discuss the ozone-CO correlation at PICO-NARE; their Figure 7 shows only a few scattered 443 points with such correlation (relatively high ozone at low CO). Consistent with other PICO- 444 NARE analyses (Honrath et. al., 2004; Lapina et al., 2006; Owen et al., 2006; and Val Martin et 445 al., 2006) the ozone variability is expected to reflect the influence of the regional photochemical 446 production and destruction in the central North Atlantic. 447 448 The evolution of NMHC ratios through photochemical processing provides a means to 449 investigate the photochemical evolution of ozone (Parrish et al., 1992; 2004). Figure 7 shows the 450 dependence of ozone concentrations on the natural logarithm of [propane]/[ethane] as the 451 indicator of the photochemical processing in each season. 452 concentrations are relatively constant with no dependence on the NMHC ratios. In spring and 453 summer ozone has higher variability, both toward higher and lower concentrations. 454 relationships in Figure 7 indicate that higher ozone levels were consistently observed in air that 455 had relatively ‘fresh’ photochemical signatures (e.g. ln [propane]/[ethane] > -2.5), and that lower 456 ozone correlated with more processed air (i.e. ln [propane]/[ethane] < -2.5). These relationships 457 suggest that in spring and summer the highest ozone concentrations are observed when air 458 masses most recently transported from continental source regions impact the site, and lower 459 concentrations are observed in air masses that have been processed for longer times in the marine 460 troposphere. Evidently the photochemical environment of aged air masses in the central North 461 Atlantic is characterized by net photochemical destruction of ozone in spring and even more 462 strongly in summer. During fall and winter ozone The 463 15 464 Table 2 compares the springtime and summertime slope of the ozone - ln [propane]/[ethane] 465 relationship found at PICO-NARE with those reported from the north temperate Pacific marine 466 boundary layer (data for fall and winter were excluded because of the weak correlation result or 467 the regression analyses (Fig. 7)). Based on these results Parrish et al. (2004) argue that the recent 468 Pacific studies (ITCT-2K2, PHOBEA, TRACE-P) find evidence for only weak net ozone 469 destruction (small positive slopes) in the more remote Pacific marine boundary layer. This weak 470 photochemical destruction is in sharp contrast with the much stronger photochemical destruction 471 indicated by a study at Point Arena from nearly two decades earlier. The exception to this 472 picture is the strong photochemical production (large negative slope) in the PEM West-B study, 473 which focused on the region of strong outflow of ozone precursor emissions from Asia to the 474 western North Pacific. Comparison of the Pacific results to those from PICO-NARE suggest that 475 spring- and summertime photochemistry in highly aged air masses more effectively destroys 476 ozone in the central North Atlantic than in the North Pacific. 477 478 3.5 Transport Event Case Study 479 480 In Figure 8a six weeks of data for four NMHC during spring 2005 are shown (note the 481 logarithmic concentration scale). NMHC concentrations are highly variable, close to 10-fold 482 increases were observed several times during this observation window. The high correlation, 483 with concurrent minima and maxima of these four individual NMHC is noteworthy. 484 amplitudes of relative mixing ratio increases are highest for the shorter-lived compounds. 485 Underneath these variable data, the springtime decline in the NMHC mixing ratios can be 486 discerned. The ln [propane]/[ethane] and ln [butane]/[ethane] analysis for the same data (Figure 487 8b) can be used to investigate the short-term changes in the inferred photochemical age of air 488 reaching PICO-NARE. Again, a high variability is found, with high ln [NMHCi]/[ethane] ratios 489 (indicating ‘fresh’, e.g. little processed air) coinciding with periods of enhanced absolute NMHC 490 mixing ratios and low ln [NMHCi]/[ethane] ratios (indicating ‘old’, e.g. well processed air) 491 coinciding with periods of low absolute NMHC mixing ratios. The 492 493 Three periods when air switched from a ‘fresh’ signature to an ‘old’ signature and back to a 494 ‘fresh’ character were subjected to a closer investigation. These three, ~ 1-2 day intervals are 16 4 495 indicated by the colored circles in Figure 8b. The corresponding data points are marked by the 496 same colors and compared with all data during this April-May period in the ln [butane]/[ethane] 497 versus ln [propane]/[ethane] plot in Figure 9. Here, we derived the line constraining the mixing 498 boundary by assuming that air, influenced with recent emissions, was not mixed with zero- 499 concentration background air (as done for Figure 5) but instead, more realistically and as 500 suggested by McKeen and Liu (1993) was mixed with background (B) air that had inferred 501 seasonal ethane, propane and butane mixing ratios of 900, 20, and 2 pptv, respectively. These 502 modifications result in the shaped form of the mixing boundary that constrains these data. 24- 503 hour [OH] was estimated at 2 x 106 molecules cm-3 (according to Spivakovsky et al 2000 et al., 504 for 800 hPa, April, 36.0oN, 27.5oW). Molar emission ratios were set to 0.63 for propane:ethane 505 and 0.35 for n-butane:ethane, which are the averages of results from Goldstein, et al (1995), and 506 Swanson, et al. (2003). The concentration change of these NMHC and their ratios can then be 507 calculated according to 508 model: 509 d N Background: 510 OH only ethane=900 Mixing and(0.1 OHh-1) and k = 0.18 x 10-12, 0.89 x 10-12 and 2.05 x 10-12 511 with K being the mixing constant propane=60 16:00 4/19 1:00respectively (Atkinson and Arey [2003] with 512 molecules cm-3s-1 for4/17 ethane, propanetoand n-butane, butane=4 4/19 17:00 to and 4/20 15:30 of hydrocarbon pairs5d 513 T=273 of this equation consideration will yield K=0.01 h-1 K). Integration 3/26 to 5/14 OH=2E6 A0 LA A LA ln ln Mixing Only B LB B0 LB 514 t Starting: kB k A OH ethane:10 propane:5 515 where, butane:4 10d 516 K AB kOH 517 , LA K k OH ethane:2.04e-13 A propane:9.77e-13 518 butane:2.29e-12 519 0d 15d and A and B are a NMHC pair. 520 -3.5 -3 -2.5 -2 -1.5 ln(propane/ethane) -1 17 -0.5 521 This equation, using the above parameters, allows assigning a hypothetical average 522 photochemical aging, respectively transport time from the occurrence of fresh emissions to 523 arrival at Pico, which is marked as the boxed numbers (in days) on the regression line of the data 524 depicted in Figure 9. Under these assumptions, the data of the three episodes marked in Figure 525 8b are defined with ages of 9-11, 15-16, and 8-10 days respectively. 526 527 Previous attempts of deriving photochemical transport times, or ‘photochemical clocks’ from 528 analysis of NMHC ratios have raised the question to what extent quantitative information from 529 this analysis can be valid. The fact that naturally the transport path, reaction history and dilution 530 of air parcels arriving at a receptor site may vary substantially, poses severe limitations on this 531 utility (ref). In the following paragraphs two synoptic transport and air parcel mixing analysis 532 approaches are applied for an evaluation of the above interpretations derived from the NMHC 533 ratio analysis. 534 535 Back trajectories for the three episodes marked periods in Figure 8b are shown in Figure 10. Air 536 sampled at Pico on 4/18 had previously been traveled across the Northern Pacific and then been 537 rapidly transported over California, the Midwestern U.S. and Northern Canada. 538 plots show that 4-6 days prior to arrival at Pico, these trajectories passed over the mid-western 539 and Western U.S.. at relatively low, e.g. 2-4 km above the surface. In contrast, air trajectories 540 arriving during 4/20 had resided within the oceanic, mid-Atlantic region for a minimum of 10 541 days. Two days later, air arriving on 4/22 had passed over Eastern Canada and the Central 542 United States 2-10 days prior. The transport altitude during this third period was lower than 543 during 4/18, indicating possibly stronger influence and more recent injections for surface 544 emissions. The altitude 545 546 The comparison of the NMHC interpretations with these trajectory analysis imply that periods 547 identified with ‘photochemically fresh’ air coincided with air transport over populated, U.S. 548 continental regions, where, most likely, an injection of recent anthropogenic emissions had 549 occurred. In contrast, the period that was identified as ‘photochemically old’ was attributed to 550 conditions where air had resided over the Atlantic ocean for an extended (> 10 days) period of 551 time. The photochemical age classification derived from our calculations presented above 18 552 qualitatively agree well with the trajectory interpretations; calculated mean photochemical 553 transport times are generally longer than inferred trajectory transport times from the most 554 recently encountered influence of expected surface influenced during the transport path. 555 556 These conclusions are further supported by the FLEXPART simulations. Retroplumes for the 557 three episodes are shown in Figure 11a, c, d are in general agreement with the trajectory analysis. 558 FLEXPART provides additional information on the spatial distribution of emission sources that 559 contributed to CO enhancements in the retroplume. The source contribution map (Figure 11b) 560 indicates that for the first event the bulk of the emissions originated over central Colorado about 561 six days prior to arrival. Significant contributions were also from the western US with age 562 ranges from 6-8 days. The contribution of individual CO enhancements (at 1-day transport tme 563 resolution) to the overall, anthropogenic CO enhancement as well as the derived average age of 564 CO from FLEXPART is shown in Figure 8c, in comparison with the NMHC time series. The 565 average age during the first event from this analysis was 7-8 days, most of the CO injections had 566 occurred 8-10 days prior to arrival. Interestingly, the transport pathway appears to be similar to 567 typical warm conveyor belt transport, a mechanism that has been observed to transport emissions 568 from the North American continent directly to the PICO-NARE station (Owen et al, 2006). 569 Indeed, there was a cold front located over this region of the U.S. on April 13-15, indicating 570 frontal transport was partially responsible for this episode. The transport pathway took the 571 emissions from the source region to high altitudes above regions with little emissions which 572 resulted in little mixing with polluted air masses of different ages and composition, giving the 573 emissions observed at the station a fairly small range of age. 574 575 In contrast, FLEXPART results show that air sampled during 4/19-20 originated over the 576 southeastern Pacific, traveled briefly over Mexico and the Gulf of Mexico, before spending 10- 577 12 days within the mid-Atlantic region, circulating around the Azores High at relatively low 578 altitudes, before arriving at the PICO-NARE station (Figure 11c). The CO time series (Figure 579 8c) shows no CO present at the station less than 10 days old, with most falling into the 15-20 day 580 old bin. While there is a strong signal from southwestern Mexico, these aged emissions come 581 from all across the eastern U.S. and the Caribbean (Figure 11d), indicating significant mixing of 582 several polluted air masses as well as dilution with unpolluted marine air. The average age of the 19 583 CO in the resulting mixture, which only has a small modeled CO enhancement, ranges from 16- 584 19 days old. 585 586 Finally, air sampled during 4/21-23 had passed over large portions of North America before 587 arriving at the PICO-NARE station (Figure 11e). The path for the retroplume is consistent with 588 another typical transport pathway to the station, export from the U.S. and subsequent transport at 589 relatively low levels in the westerly wind (Owen, et al, 2006). The height of the plume during 590 transit over the U.S. was relatively low, generally less than 3 km, and remained low during 591 transport over the Atlantic to the station. The contribution map indicates that sources across the 592 eastern U.S. were responsible for the enhancements observed at this time (Figure 11f), with a 593 wide range of CO ages present. The resulting average age of CO for this episode ranges from 6- 594 9 days, though the individual ages range widely, from 4-15 days old (Figure 8c). Both the 595 retroplumes and distribution of CO ages indicate the mixing of many air masses of relatively 596 fresh emissions (particularly compared to the air sampled during 4/19-20). 597 598 In general, FLEXPART simulates not only the timing of CO enhancements well, but also 599 appears to simulate the relative ages of emissions. The average age appears to be the best value 600 to compare with the “photochemical age” derived from the HC ratios. Conclusively, both the 601 back-trajectory analysis and FLEXPART calculations yield reasonable confirmation of the 602 aforementioned interpretations from observed concentrations of NMHC and their photochemical 603 processing and mixing during transport. 604 605 Can we improve/expand last paragraph? Which comparison gives better agreement? Why? 606 What are limitations of photochemical clock? How can the photochemical clock possibly be 607 improved? 608 609 4. Summary and Conclusions 610 611 Air sampled at PICO-NARE shows high variability in NMHC and their ratios during all times of 612 the year. This observation is indicative of the variable atmospheric transport conditions that 613 bring in air with variable flow and with much different origin and photochemical history. 20 614 Overall, concentrations of NMHC at PICO-NARE are higher than at MLO, which reflects the 615 higher influence of the adjacent continents to air composition in the central Atlantic region in 616 comparison to the Northern Mid-Pacific. 617 618 Short-chain NMHC remain elevated in air plumes that have been influenced by either 619 anthropogenic injections or biomass burning after time scales in excess of 1 week during their 620 transport to the PICO-NARE station. Isoprene data convincingly describe summertime (mostly 621 buoyant) upslope flow occurrences. Isoprene was found as the best of all chemical tracers to 622 identify upslope flow. 623 624 A good correlation was determined between seasonally differentiated NMHC variability and the 625 NMHC OH lifetimes. Regression analysis of the lnx=A-b relationship for these data yields a b- 626 value of -0.44, which confirms the remote, marine island character of the Pico site and the lack 627 of major local influences on NMHC levels. 628 629 Summertime ozone/(ln [propane]/[ethane]) correlations show higher variability, indicating more 630 variability in photochemical conditions than during wintertime. Net ozone destruction typically 631 only occurs, after photochemical processing has reduced ln [propane]/[ethane] to values < -2.5; a 632 conditions only observed during the summer. 633 634 The reasonable agreement that was found between indirectly derived photochemical ages of 635 NMHC in air plumes sampled at Pico and back-trajectory and FLEXPART analysis suggests that 636 assumptions that went into the model calculations were reasonable and that interpretations of 637 NMHC ratios provide a meaningfull tool for deciphering photochemical age and transport of air 638 sampled at Pico. 639 640 Acknowledgments 641 642 We thank P. Goldan, NOAA Aeronomy Laboratory, Boulder, CO for the reference analysis of 643 the primary NMHC standard prior and after its use at Pico, M. Dziobak and M. Val Martin, 644 Michigan Technological University, for GC instrument maintenance tasks at the Pico, D. 21 645 Henriques, Institute of Meteorology, Ponta Delgada, Portugal for retrieving the ECMWF data 646 used in this work, Andreas Stohl, Norsk Institutt for Luftforskning (NILU), Kjeller, Norway for 647 providing and assisting in running the FLEXPART mode, the Data Support Section of NCAR’s 648 Scientific Computing Division for making the NCEP FNL analyses available for download and 649 T. Jobson, Washington State University, for the Fraserdale data. This research was funded by a 650 grant from the NOAA Office of Global Programs (award # NA03OAR4310072). 651 652 653 654 655 References Atkinson R. and J. Arey (2003), Atmospheric degradation of volatile organic compounds, Chemical Reviews, 103, 4605-4638 656 657 Blake D.R. and F.S. Rowland (1995) Urban leakage of liquefied petroleum gas and its impact on Mexico City air quality. Science 269, 953-956. 658 659 Bottenheim J.W. and M.F. Shepherd (1995) C2-C6 hydrocarbon measurements at four rural locations across Canada. Atmos. Environ. 29, 647-664. 660 661 Bottenheim J.W., P.C. Brickell, T.F. Dann, D.K. Wang, F. Hopper, A.J. Gallant, K.G. Anlauf and H.A. Wiebe (1997) Non-methane hydrocarbons and CO during Pacific ’93. Atmos. Environ. 14, 2079-2087. 662 Danielsen, E. F., R. S. Hipskind, S. E. Gaines, G. W. Sachse, G. L. Gregory, and G. F. 663 664 Hill, Three-dimensional analysis of potential vorticity associated with tropopause folds and observed variations of ozone and carbon monoxide, J. Geophys. Res., 92, 2103-2111, 1987. 665 666 667 668 669 De Gouw J.A., Cooper O.R., Warneke C., Hudson P.K., Fehsenfeld F.C., Holloway J.S., Hübler G., Nicks D.K. Jr., Nowak J.B., Parrish D.D., Ryerson T.B., Atlas E.L., Donnelly S.G., Schauffler S.M., Stroud V., Johnson K., Carmichael G.R., and Streets D.G. (2004) Chemical composition of air masses transported from Asia to the U.S. West Coast during ITCT 2K2: Fossil fuel combustion versus biomassburning signatures. J. Geophys. Res. 106, D23S20, doi: 10.1029/2003JD004202. 670 671 672 Draxler R. and Rolph G. (2003) HYSPLIT4 (Hybrid Single-Particle Lagrangian Integrated Trajectory) model. Air Resour. Lab., Natl. Oceanic and Atmos. Admin., Silver Spring, Md. (available at http://www.arl.noaa.gov/ready/hysplit4.html). 673 674 ECMWF (2005), Users guide to ECMWF products 4.0, Tech. Rep. Meteorological Bulletin M3.2, European Center for Medium-Range Weather Forcasts (ECMWF), Reading, UK. 675 676 Ehhalt D.H., F. Rohrer, A. Wahner, M.J. Prather and D.R. Blake (1998) On the use of hydrocarbons for the determination of tropospheric OH concentrations. J. Geophys. Res. 103, 18981-18,997. 677 678 679 Gautrois M., T. Brauers, R. Koppmann, F. Rohrer, O. Stein and J. Rudolph (2003) Seasonal variability and trends of volatile organic compounds in the lower polar troposphere. J. Geophys. Res. 108, 4393, doi:10.1029/2002JD002765. 680 681 682 Goldstein A.H., S.C. Wofsy, and C.M. Spivakovsky (1995) Seasonal variations of nonmethane hydrocarbons in rural New England: constraints on OH concentrations in northern midlatitudes. J. Geophys. Res. 100, 21023-21033. 683 684 685 Greenberg, J.P., D. Helmig, and P.R. Zimmerman, Seasonal measurements of nonmethane hydrocarbons and carbon monoxide at the Mauna Loa Observatory during the Mauna Loa Observatory Photochemical Experiment 2 (1996), J. Geophys. Res., 101, 14581-14598. 22 686 687 688 689 Honrath R.E., R.C. Owen, M. Val Martin, J.S. Reid, K. Lapina, P. Fialho, M.P. Dziobak, J. Kleissl, and D.L. Westphal (2004), Regional and hemispheric impacts of anthropogenic and biomass burning emissions on summertime CO and O3 in the North Atlantic lower free troposphere, J. Geophys. Res., 109, D24310, doi;10.1029/2004JD005147. 690 691 692 Fehsenfeld, F., Calvert, J., Fall, R., Goldan, P., Guenther, A. B., Hewitt, C. N., Lamb, B., Shaw, L., Trainer, M., Westberg, H., and Zimmerman, P. (1992). “Emissions of volatile organic compounds from vegetation and the implications for atmospheric chemistry.” Global Biogeochem. Cycles, 6, 389-430. 693 694 Fialho P., Freitas M.C., Barata F., Vieira B., Hansen A.D.A. and Honrath R.E. (2006) The Aethalometer calibration and determination of iron concentration in dust aerosols. J. Aerosol Sci., in press. 695 696 697 Finlayson-Pitts B.J. (1993) Comment on “Indication of photochemical histories of Pacific air masses from measurements of atmospheric trace species at Point Arena, California”, by D.D. Parrish et al. J. Geophys. Res. 98, 14991-14993. 698 699 Jobson B.T., Wu Z., Niki H. and Barrie L.A. (1994) Seasonal trends of isoprene, C2-C5 alkanes, and acetylene at a remote boreal site in Canada. J. Geophys. Res. 99, 1589-1599. 700 701 702 Jobson B.T., D.D. Parrish, P. Goldan, W. Kuster, F.C. Fehsenfeld, D.R. Blake, N.J. Blake and H. Niki (1998) Spatial and temporal variability of nonmethane hydrocarbon mixing ratios and their relation to photochemical lifetime. J. Geophys. Res. 103, 13557-13567. 703 704 705 Jobson B.T., S.A. McKeen, D.D. Parrish, F.C. Fehsenfeld, D.R. Blake, A.H. Goldstein, S.M. Schauffler, and J.W. Elkins (1999) Trace gas mixing ratio variability versus lifetime in the troposphere and stratosphere: Observations. J. Geophys. Res. 104, 16091-16113. 706 707 708 Jobson B.T., C.M. Berkowitz, W.C. Kuster, P.D. Goldan, E.J. Williams, F.C. Fehsenfeld, E.C. Apel, T. Karl, W.A. Lonneman and D. Riemer (2004) Hydrocarbon source signatures in Houston, Texas: Influence of the petrochemical industry. J. Geophys. Res. 109, D24305, doi:10.1029/2004JD004887. 709 710 711 Kleissl J., R.E. Honrath, M.P. Dziobak, D. Tanner, M. Val Martin, R.C. Owen, and D. Helmig (2006) The occurrence of upslope flows at the Pico mountaintop observatory: A case study of orographic flows on a small, volcanic island. J. Geophys. Res., submitted for publication. 712 713 McKeen S.A. and S.C. Liu (1993) Hydrocarbon ratios and photochemical history of air masses. Geophys. Res. Let. 20, 2363-2366. 714 715 716 Olivier, J.G.J. and J.J.M. Berdowski (2001) Global emissions sources and sinks. In: Berdowski, J., Guicherit, R. and B.J. Heij (eds.) "The Climate System", pp. 33-78. A.A. Balkema Publishers/Swets & Zeitlinger Publishers, Lisse, The Netherlands. ISBN 90 5809 255 0. 717 718 Owen, R., O. Cooper, A. Stohl, and R. Honrath (2006), An analysis of transport mechanisms of North American emissions to the Central North Atlantic, J. Geophys. Res., doi:1029/2006JD007062, in press. 719 720 721 722 Parrish D.D., C.J. Hahn, E.J. Williams, R.B. Norton, F.C. Fehsenfeld, H.B. Singh, J.D. Shetter, B.W. Gandrud, and B.A. Ridley (1992) Indications of photochemical histories of Pacific air masses from measurements of atmospheric trace species at Point Arena, California. J. Geophys. Res. 97, 15,88315,901. 723 724 725 726 Parrish D.D., M. Trainer, V. Young, P.D. Goldan, W.C. Kuster, B.T. Jobson, F.C. Fehsenfeld, W.A. Lonneman, R.D. Zika, C.T. Farmer, D.D. Riemer and M.O. Rodgers (1998) Internal consistency tests for evaluation of measurements of anthropogenic hydrocarbons in the troposphere. J. Geophys. Res. 103, 22339-22359. 727 728 Parrish D.D., E.J. Dunlea, E.L. Atlas, S. Schauffler, S. Donnelly, V. Stroud, A.H. Goldstein, D.B. Millet, M. McKay, D.A. Jaffe, H.U. Price, P.G. Hess, F. Flocke, and J.M. Roberts (2004) Changes in the 23 729 730 photochemical environment of the temperate North Pacific troposphere in response to increased Asian emissions. J. Geophys. Res. 109, D23S18, doi:10.1029/2004JD004978. 731 732 733 Penkett S.A., N.J. Blake, P. Lightman, A.R.W. Marsh, P. Awyl, and G. Butcher (1993) The seasonal variation of nonmethane hydrocarbons in the free troposphere over the North Atlantic Ocean: Possible evidence for extensive reaction of hydrocarbons with the nitrate radical. J. Geophys. Res. 98, 2865-2885. 734 735 Rudolph J. (1995) The tropospheric distribution and budget of ethane. J. Geophys. Res. 100, 1136911381. 736 737 Stohl A., Hittenberger M. and Wotawa G. (1998) Validation of the Lagrangian particle dispersion model FLEXPART against large scale tracer experiments. Atmos. Environ. 32, 4245-4264. 738 739 Stohl A., Forster C., Frank A., Seibert P., and Wotawa G. (2005) Technical Note : The Lagrangian particle dispersion model FLEXPART version 6.2. Atmos. Chem. Phys. 5, 2461-2474. 740 741 Stohl, A., M. Hittenberger, and G. Wotawa (1998), Validation of the Lagrangian particle dispersion model FLEXPART against large scale tracer experiment data, Atmos. Environ., 32, 4245-4264. 742 743 Stohl, A., and D. J. Thomson (1999), A density correction for Lagrangian particle dispersion models, Boundary-Layer Meteorol., 90, 155-167. 744 745 746 Swanson, A.L., et al. (2003), Seasonal varioations of C2-C4 nonmethane hydrocarbons and C1-C4 alkyl nitrates at the Summit research station in Greenland, J. Geophys. Res., 108, 4065, doi:4010.1029/2001JD001445. 747 748 749 Tanner D., D. Helmig, J. Hueber and P. Goldan (2006) Gas chromatography system for the automated, unattended, and cryogen-free monitoring of C2 to C6 non-methane hydrocarbons in the remote troposphere. J. Chrom., 1111, 76-88. 750 751 752 Val Martin M., R.E. Honrath, R.C. Owen, G. Pfister, P. Fialho and F. Barata, Significant enhancements of nitrogen oxides, black carbon and ozone in the North Atlantic free troposphere resulting from North American boreal wildfires. J. Geophys. Res., submitted for publication. 753 754 755 Williams J., V. Gros, B. Bonsang and V. Kazan (2001) HO cycle in 1997 and 1998 over the southern Indian Ocean derived from CO, radon, and hydrocarbon measurements made at Amsterdam Island. J. Geophys. Res. 106, 12719-12725. 756 757 758 759 Williams J., H. Fisher, G.W. Harris, P.J. Crutzen, P. Hoor, A. Hansel, R. Holzinger, C. Warneke, W. Lindinger, B.. Scheeren and J. Lelieveld (2000) Variability-lifetime relationship for organic trace gases: A novel aid to compound identification and estimation of HO concentrations. J. Geophys. Res. 105, 20473-20486. 24 Figure 1 Whisker plots of monthly data for ethane, propane, i-butane, n-butane, i-pentane and pentane. The 5, 25, 50, 75, 95 percentiles are indicated by the horizontal lines of each box, the vertical lines extend to the minimum and maximum observed values. The width of the box indicates the time period over which data was acquired for a given month. The vertical dotted lines show the windows that were applied in the seasonal (winter, spring, summer, fall) analysis. Data from a remote boreal forest site in Canada (Jobson et al., 1994) and 25, 50, and 75 percentile data from Mauna Loa Observatory (Greenberg et al., 1996) were added for comparison. 25 Figure 2 Cumulative distributions of NMHC at the PICO-NARE station during the four measurement seasons. 26 Figure 3 The standard deviation of the natural logarithm of the NMHC mixing ratio during the four seasons at its seasonal OH lifetime. 27 the seasonal data distribution. 28 Figure 5 Mixing ratio of i-butane versus n-butane (left) and ratio of i-butane/n-butane versus n-butane (right) in the upper graphs and of n-pentane versus i-pentane (left) and ratio of n-pentane/ipentane versus i-pentane (right) in the lower graphs The error bars in the right graph show the standard deviation of the data within 10-percentile bins of the data distribution. Dotted lines illustrate the estimated uncertainties in the measurement. REMOVE dotted uncertainty ranges? Figure 6 Relationship between the natural logarithms of [n-buante]/[ethane] versus [propane]/[ethane] for the fall, summer, spring and summer data as defined in Fig. 1. 29 Figure 7 Ozone in relation to the natural logarithm of [propane]/[ethane] (indicating degree of photochemical processing) during the fall, winter, spring and summer. The lines indicate the linear least-squares fits to the log-transformed data, and their slopes with 95% confidence limits and correlation coefficients are annotated. 30 A B 4/17 16:00 to 4/19 1:00 4/21 5:00 to 4/23 15:30 4/19 17:00 to 4/20 15:30 31 C Figure 8 Mixing ratios of four NMHC (A) and the natural logarithms of [propane]/[ethane] and [butane]/[ethane] for 21 days in spring 2005 (B). Panel (C) shows for the same time window the Flexpart calculations for the contribution of emissions of CO over North America on the CO enhancement in the retroplume arriving at Pico. The derived average CO enhancement transport time is shown on the secondary y-axis. . Still need missing May data. Try to redo figure c in Excel so that format matches panels a and b? please change y-axis title to CO (ppbv). 32 Figure 9 Distribution of the marked data points in Fig. 8 (graph B) in the ln-ln photochemical age/dilution analysis plot in comparison to all spring 2005 data (top). A photochemical age scale was calculated using variables described in the text. Resulting ages (in days of transport) are indicated in the diamonds on the regression line fit. 33 Figure 10 Geographical and altitude back trajectory analysis of the events marked in the Fig. 8b. The solid lines illustrate back trajectories (every 6 hours) arriving at six grid points surrounding the station for 4/18, 00.00 UTC. The dotted lines show the corresponding analysis for 4/20 00.00 UTC and the dashed back trajectories are for arrival at 4/22 00:00 UTC. Dot marks on the trajectory lines indicate 2-day transport distances. Right now this is a composite of the same figure cut and pasted twice, zoomed to different sizes. Can we make this one nice figure at a somewhat improved resolution? 34 Figure 11 Results for retroplumes initiated at 00-03 on July 18 (A&B) July 20 (C&D) and July 22 (E&F). The left column shows the total column (0-10km) residence times. The right column shows the foot print layer (0-300m) response to emissions sources (residence time folded with emission strength). Colors are logarimithically scaled (100-1%) according to the maximum value for each plot type (14000 seconds for total residence time, 4700 grams of CO for foot print layer response), as shown by the scale on the bottom. 35 Table 1 36