Você está na página 1de 212

Primera Edicin Diseo y Edicin: Collage Creativo Fotografa: Durga Archivo Digital de Imgenes www.durga.com.

mx DR Instituto Nacional de Ecologa (INE-Semarnat) Perifrico Sur No. 5000, Col. Insurgentes Cuicuilco, C.P. 04530, Mxico, D.F. www.ine.gob.mx La publicacin de esta tesis estuvo bajo el patrocinio de la Agencia de Cooperacin Internacional del Gobierno de Japn (JICA) ISBN: 968-817-563-3 Impreso y hecho en Mxico.

Formation and Transformation Mechanisms of Particulate Matter Under Ten Micromilimeters (PM 10 ) and Ozone (O 3 ) in the Mexico City Metropolitan Area and the Greater Manchester Area

Por qu publicar este trabajo?


De manera muy especial y en calidad de edicin emblemtica, el Instituto Nacional de Ecologa (INE) a travs del Centro Nacional de Investigacin y Capacitacin Ambiental (CENICA) publica, como un modesto reconocimiento pstumo al Dr. Gabriel de Icaza del Ro, su tesis doctoral Formation and Transformation Mechanisms of Particulate Matter Under Ten Micrometers (PM10) and Ozone (O3) in the Mexico City Metropolitan Area and the Greater Manchester Area, realizada en la Universidad de Manchester, y fruto del esfuerzo ejemplar que desarroll durante sus ltimos aos de vida. La realizacin de esta tesis fue posible gracias al patrocinio de la Agencia de Cooperacin Internacional del gobierno del Japn (JICA por sus siglas en ingls), organizacin que apoy con tcnicos sobresalientes y con generoso financiamiento el desarrollo del CENICA y el trabajo de Gabriel de Icaza. Aunque, trgicamente, Gabriel no pudo vivir lo suficiente como para convertir su trabajo de tesis en publicacin formal, JICA decidi proponer su publicacin y como tantas veces lo ha hecho proporcion el apoyo financiero para el trabajo editorial. La publicacin emblemtica de esta tesis se hace como una celebracin a la amistad entre Mxico y Japn, y como una conmemoracin a Gabriel de Icaza, uno de los artfices de esa slida amistad. El Dr. Gabriel de Icaza del Ro lleg al CENICA en 1999, despus de haber realizado estudios de maestra en el Instituto de Ciencia y Tecnologa de la Universidad de Manchester, Inglaterra, y de haber obtenido el grado de Doctor en Filosofa. Para esa poca ya saba que su salud estaba fuertemente quebrantada, no obstante, su alegra, vitalidad y capacidad siempre estuvieron presentes al incorporarse a las filas de la investigacin en el rea de la contaminacin atmosfrica, que por ese entonces empezaba a fortalecerse al interior del Centro y a la que contribuy de manera determinante. En efecto, durante su efmera pero slida presencia por los pasillos del CENICA, y del brazo con el personal bajo su coordinacin, imprimi nuevos bros y diversific los caminos para el estudio y anlisis sobre uno de los problemas ambientales que ms apremia a la humanidad en su conjunto, vivir en un ambiente sano y respirar un aire limpio, al que todos tenemos derecho. Inteligente y bien dispuesto, organizaba y propona rumbos de trabajo, a la par que atenda con esmero el desarrollo profesional de quienes estaban bajo su direccin. Sociable, ameno y con un estado de nimo inquebrantable siempre tena a la mano una frase de aliento y comprensin durante las faenas, a veces ridas, que orientara el trabajo an inconcluso del proyecto en puerta. Con su desaparicin prematura el Dr. Gabriel de Icaza del Ro dej un hueco difcil de llenar en nuestras filas, ejemplo claro de lo que puede hacer la inteligencia, la sensibilidad y el coraje en aras de la libertad y de la vida. Sirva pues esta publicacin como un pequeo tributo y en recuerdo al legado que dej entre nosotros. Vctor Javier Gutirrez Avedoy El Director General del Centro Nacional de Investigacin y Capacitacin Ambiental
3

Semblanza del autor


Ama y haz lo que quieras (San Agustn) La vida, esa amorosa amiga que lo arrancara de sus brazos en la euforia de su pasin, fue a la que le dedic lo mejor de sus horas para dejarnos en estos sesudos anlisis, como gotas pesadas y ricas de esfuerzo, su aportacin a la ciencia del medio ambiente. Inici los estudios comparados en la materia analizando dos ciudades para l prximas a su espritu, Mxico y Manchester. La que lo vio nacer y en la que floreci su intelecto y le aportara las herramientas que aqu, en su tiempo y en su obra, hoy habrn de servir al futuro para una vida mejor. Gabriel de Icaza del Ro escribi el primero de febrero de 1993 en Pasadena, California, en un ensayo autobiogrfico lo siguiente: Determinacin, disciplina y esfuerzo son las herramientas claves que he tratado de emplear para lograr mis objetivos. En mi vida he sido expuesto a dos culturas diferentes y he aprendido a lidiar con ambas. He sido enfrentado al reto de aprender otro idioma y tambin a adaptarme a otra cultura. Esto ha dado por resultado mi autodisciplina y mi automotivacin. Esta experiencia bicultural me hizo reconocer que cuando concentro mi mente en algo, lo puedo lograr. Nac en la Ciudad de Mxico el 19 de mayo de 1970 en donde viv durante los primeros trece aos de mi vida. Termin la primaria en una escuela privada bilinge llamada Queen Elizabeth School. Despus me mud a los Estados Unidos de Amrica y empec a estudiar en la escuela catlica Daniel Murphy High School. Despus me cambie y me gradu en el Pasadena High School en donde conoc los programas de estudio del Pasadena City College. Eleg esta Institucin por su compromiso con la excelencia en las reas de matemticas y ciencias; una vez cubierto el ciclo escolar y algunos cursos extras, ingres a la University of California, Los Angeles, (UCLA) en donde me gradu en Geografa / Ecosistemas e Ingeniera Ambiental. Decid esta rea porque me preparara tanto en los aspectos tcnicos como en los sociales de las ciencias ambientales, que pienso son necesarias para completar una formacin profesional integral. En m, siempre ha existido una curiosidad por explorar los problemas y las soluciones que lidian con ambos: la naturaleza y el hombre. Este inters me llevo a explorar casi cualquier cosa que caa en mis manos e incluso en ocasiones me meti en dificultades, pero siempre logr extraer experiencias valiosas. Yo creo que la mejor manera de satisfacer esta curiosidad es continuar mi educacin como ingeniero ambiental al nivel de postgrado. Yo creo que en los desafos esta la ruta del progreso, y en el progreso est la auto superacin.

En mi vida he tenido que enfrentar tres retos fundamentales: primero, yo nac con una discapacidad fsica que en mis primeros aos requiri una gran cantidad de energa para superarla. Esto signific un reto tcnico porque necesitaba una prtesis ortopdica que tena que ser diseada especialmente para mi por lo raro de mi problema, y porque mi inters por llevar una vida normal hacia necesario optimizar funcin, peso y cosmtica, por lo que tena que trabajar junto con los tcnicos en los procesos de diseo y construccin. Segundo, me tuve que adaptar a una cultura diferente al tiempo que atravesaba por mi adolescencia. La tarea bsica consisti en llenar la brecha entre el ingls conversacional que haba aprendido en Mxico y el ingls acadmico que necesitaba para tener xito en la escuela. El tercero, que an esta en evolucin, es continuar mi educacin. Todos los logros que me he ganado hasta ahora, y los retos que he enfrentado, cumplen con su propsito; si tengo acceso a la mejor educacin. Esto me capacitar para cumplir con mis objetivos vocacionales y me dar la garanta de realizar mi potencial como profesional. Me siento afortunado del hecho de tener el deseo de contribuir no slo a la comunidad en que nac y crec, sino tambin de enriquecer la vida de muchas otras comunidades. El mejoramiento de la situacin ambiental es un factor positivo; no slo para un lugar sino que para todo el planeta. Poco despus el Consejo Nacional de Ciencia y Tecnologa de Mxico otorg a Gabriel una beca para continuar sus estudios en la University of Manchester Institute of Science and Technology, donde obtuvo el grado de Master in Science in the Faculty of Technology en 1995. En junio de 1999 el Departamento de Fsica de la misma institucin le otorg el grado de Doctor of Philosophy por su contribucin a la ciencia e inmediatamente se incorpor al Instituto Nacional de Ecologa en el Centro Nacional de Investigacin y Capacitacin Ambiental, como Subdirector de Investigacin de la Contaminacin Atmosfrica. Jos Antonio de Icaza Gmez Lorenza del Ro Caedo Padres Esteban de Icaza del Ro Lorenza de Icaza del Ro Hermanos

Formation and Transformation Mechanisms of Particulate Matter Under Ten Micrometers (PM10) and Ozone (O3) in the Mexico City Metropolitan Area and the Greater Manchester Area

A thesis submitted to the University of Manchester Institute of Science and Technology for the degree of Doctor of Philosophy June 1999 Gabriel de Icaza del Ro Department of Physics I would like to dedicate this thesis to God all mighty and the Virgin of Guadalupe, to my beloved parents and to my brother and sister. Special gratitude is expressed to Alexandra, whose patience and encouragement were always present. Beverly, thank you for checking my English. Most importantly, I thank my supervisor Tom Choularton for giving me endless advice, for his patience and for his fruitful comments. All the work undertaken for this thesis was made possible thanks to a financial grant from CONACYT. No portion of the work referred to in the present thesis has been submitted in support of an application for another degree or qualification at this or any other university, or any other institution of learning.

Table of Contents
Table of Figures Abstract Introduction Boundary layer Meteorology 1 1.1 1.11 1.2 1.3 1.31 1.311 1.312 1.32 1.33 1.331 1.332 1.333 1.4 1.41 1.42 1.43 1.5 Introduction Radiation Radiation Balance Temperature and Pressure Boundary Layer Forcings Wind General Circulation of the Atmosphere Surface Wind Atmospheric Stability Boundary Layer Depth and Structure Mixed Layer Residual Layer Stable Layer Urban Boundary Layer Urban Heat Island Wind Urban Boundary Layer Characteristics Conclusion

9 13 17 19

23 23 25 26 29 29 30 31 33 34 35 35 36 37 38 40 41 42

Atmospheric Aerosols 2 2.1 2.2 2.21 2.22 2.23 2.3 2.4 2.5 Introduction Atmospheric Aerosols Formation of Aerosols Nucleation Coagulation Aqueous Phase Reaction Mechanisms Deleterious Effects Removal Mechanisms Conclusion 45 45 47 47 50 50 53 54 57

Ozone Formation 3 3.1 3.2 3.3 Introduction Historical Perspective Basic Photochemistry Ozone Formation
9

61 61 62 64

3.31 3.31 3.32 3.33 3.331 3.34 3.35 3.4

Basic Photochemical Reactions Sources of Radical Species Photochemistry of the Background Troposphere Photochemistry of the Urban Troposphere Chemistry of NOx and Organic Compounds Tropospheric Sinks of Ozone and Its Precursors VOC to NOx Ratio (VOC/NOx) Conclusion

64 66 68 69 70 72 75 76

Metropolitan Areas 4 4.1 4.11 4.12 4.13 4.14 4.141 4.142 4.1421 4.1422 4.1423 4.1424 4.145 4.2 4.21 4.22 4.23 4.24 4.241 4.3 Introduction Mexico City Metropolitan Area (MCMA) Geographic Description Factors Affecting Pollution Levels Environmental Controls Automated Monitoring Network of the MCMA (RAMA) Development of the Monitoring Network Atmospheric Monitoring System Automatic Network Manual Network Meteorological Network Supporting Units Metropolitan Index Greater Manchester Area (GMA) Geographical Description Factors Affecting Pollution Levels Environmental Controls Automated Urban Network of the GMA Monitoring Stations in the GMA Conclusion 81 81 82 82 83 85 86 87 87 88 88 88 89 91 91 92 94 95 96 97

Results from the MCMA 5 5. 1 5.11 5.12 5.13 5.15 5.2 5.3 5.4 5.5 Introduction Time Series PM10 Time Series Ozone Time Series Precursor Species Meteorology Time Series Diurnal Analyses Statistical Analysis Wind Analyses Conclusion 101 102 102 107 110 115 120 123 129 132

10

Discussion of the Results from the MCMA 6 6.1 6.11 6.12 6.13 6.2 6.21 6.22 6.3 6.4 6.5 6.6 Introduction Time Series PM10 O3 Meteorology Pollutant Associations PM10 O3 Diurnal Analyses Statistical Analyses Wind Analyses Conclusion 135 135 135 138 139 140 141 142 143 145 148 150

Results from the GMA 7.1 7.2 7.21 7.22 7.23 7.24 7.25 7.26 7.3 7.4 7.5 7.6 Introduction Time Series PM10 Time Series SO2 Time Series Ozone Time Series NOx (NO and NO2) Time Series Meteorological Parameters Meteorological Parameters Rainfall Time Series Diurnal Analyses Statistical Analyses Wind Analyses Conclusion 155 156 156 157 159 160 163 165 166 173 179 182

Discussion of the Results from the GMA 8 8.1 8.11 8.12 8.13 8.2 8.21 8.22 8.3 8.4 8.5 8.6 Introduction Time Series Analysis PM10 O3 Meteorology Pollutant Associations PM10 Associations O3 Associations Diurnal Analyses Statistical Analyses Wind Analyses Conclusion 185 185 186 187 188 189 190 192 193 195 196 197

11

Comparative Analysis 9 9.1 9.2 9.21 9.22 9.3 Introduction Meteorology Pollutant Associations PM10 O3 Conclusion 201 201 206 206 209 211

Conclusion Conclusion Urban Areas Meteorology Pollutants Further Research References References 223 215 215 216 217 219

12

Table of Figures
Table 1.1 Table 1.2 Table 1.3 Figure 1.1 Energy Received from the Sun, 100 Units 25 Gains and Loses by the Earth 26 Gains and Loses of Energy by the Atmosphere 26 (a) and 1.1(b) Variation of wind direction with altitude (a) Balance in the upper atmosphere, with geostrophic wind CF = PGF. (b) Balance in the surface layer CF +F = PGF 32 Table 1.4 Roughness Lengths for various surfaces 32 Figure 1.3 Profiles of mean virtual potential temperature, V, 37 Figure 1.4 Urban Boundary Layer Structure 38 Figure 1.5 Diurnal Variation in the Urban and Rural boundary Layers 39 Table 1.5 Typical roughness length (Z0) of urban terrain 40 Figure 1.6 Potential Temperature Profiles 41 Table 2.1 PM2.5 and PM10 Chemical Composition in Burbank, California1987 53 Table 3.1 Typical Atmospheric Wavelengths and Associated Energies 63 Figure 3.1 Ozone Isopleth (adapted from Seinfeld, 1986) 76 Figure 4.1 Pollution Inventory by Sector, MCMA 1994 (DDF, 1996) 84 Table 4.1 Mexican Standards 85 Table 4.2 Size Comparison of Monitoring Network over the World 87 Table 4.3 Quantity of Monitors and Mode of Analysis (RAMA) 87 Table 4.4 Meteorological Parameters and Mode of Analysis 88 Figure 4.2 Percentage Distribution of Faults (RAMA) 89 Table 4.5 Representative Values of the IMECA 90 Table 4.6 Algorithms Used to Calculate the IMECAs 90 Table 4.7 Population in Manchester and the GMA 93 Table 4.8 Local governments in the GMA, 1991 93 Table 4.9 National Air Quality Strategy, World Health Organisation Standards 95 Table 4.10 GMA Monitoring Stations 97 Figure 5.1 Map of the MCMA and the location of the Stations used in the study 101 Figure 5.2 PM10 Time Series 1994-1996 102 Figure 5.3 PM10 Yearly Trend excluding XAL 103 Figure 5.4 PM10 Monthly Averages 1994 103 Figure 5.5 PM10 Monthly Averages 1995 104 Figure 5.6 PM10 Monthly Averages 1996 104 Figure 5.7 Photograph of the Xalostoc Station in the Northeast 105 Table 5.1 PM10 vs. Wind Speed, 1994 105 Figure 5.8 PM10 and Wind Speed Time Series, 1994-96 106 Figure 5.9 PM10, Temperature and Relative Humidity, 1994-96 106 Figure 5.10 O3 Time Series 1994-1996 107 Figure 5.11 O3 Yearly Trend 108 Figure 5.12 O3 Monthly Average 1994 108 Figure 5.13 O3 Monthly Average 1995 109 Figure 5.14 O3 Monthly Average 1996 109 Figure 5.15 SO2 Time Series 1994-1996 110
13

Figure 5.16 SO2 Yearly Trends Figure 5.17 NO2 Time Series 1994-1996 Figure 5.18 NO2 Yearly Trend Figure 5.19 NO Time Series 1994-1996 Figure 5.20 NO Yearly Trend Figure 5. 21 PM10, SO2 and NO2 Time Series, 1994-96 Figure 5.22 O3, NO2 and NO, 1994-96 Figure 5.23 Temperature Time Series 1994-1996 Figure 5.24 Temperature Yearly Trend Figure 5.25 RH Time Series 1994-1996 Figure 5.26 RH Yearly Trend Figure 5.27 Wind Speed Time Series 1994-1996 Figure 5.28 Wind Speed Yearly Average Table 5.2 Rainfall Data, 1994-96 Figure 5.29 Precipitation Time Series, 1994-96 Figure 5.30 Monthly Precipitation 1994-96 Figure 5.31 Cumulative Frequency, 1994-96 Table 5.3 Emission Inventory of the MCMA by Sector (Quantities in Thousands per annum) Figure 5.32 PM10 Diurnal Variation Figure 5.33 SO2 Diurnal Variation Figure 5.34 O3 Diurnal Variation Figure 5.35 NO2 Diurnal Variation Figure 5.36 NO Diurnal Variation Table 5.4 Statistical Analyses and Durbin Watson Test Figure 5.37 PM10 vs. SO2 Figure 5.38 PM10 vs. NO2 Figure 5.39 PM10 vs. Temperature Figure 5.40 PM10 vs. Relative Humidity Figure 5.41 O3 vs. NO2, 1996 Figure 5.42 O3 vs. NO, 1996 Figure 5.43 O3 vs. Temperature Figure 5.44 O3 vs. Relative Humidity Figure 5.45 Prevailing Wind Direction 1994-96 Figure 5.46 Wind Frequency, 1994-96 Figure 5.47 PM10 vs. Wind Direction, 1994-96 Figure 5.48 SO2 vs. Wind Direction, 1994-96 Figure 5.49 NO2 vs. Wind Direction, 1994-96 Figure 5.50 O3 vs. Wind Direction, 1994-96 Figure 5.51 NO vs. Wind Direction, 1994-96 Table 6.1 PM10 and Wind Speed, 1994 Table 6.2 Pollutant Trajectory during 1994-96 in three stations Table 7.1 Monitoring Stations and Parameters Monitored Figure 7.1 Map of the GMA with the Monitoring Stations (Adapted from Barlow, 1995)

111 112 112 113 113 114 114 115 115 116 116 117 117 118 119 119 120 120 121 121 122 122 123 124 125 125 126 126 127 127 128 128 129 130 130 131 131 132 132 137 149 149 155

14

Figure 7.2 Figure 7.3 Figure 7.4 Figure 7.5 Figure 7.6 Figure 7.7 Figure 7.8 Figure 7.9 Figure 7.10 Figure 7.11 Figure 7.12 Figure 7.13 Figure 7.14 Figure 7.15 Figure 7.16 Figure 7.17 Figure 7.18 Figure 7.19 Figure 7.20 Figure 7.21 Figure 7.22 Figure 7.23 Figure 7.24 Figure 7.25 Figure 7.26 Figure 7.27 Figure 7.28 Table 7.4 Figure 7.29 Figure 7.30 Figure 7.31 Figure 7.32 Figure 7.33 Figure 7.34 Figure 7.35 Figure 7.36 Figure 7.37 Figure 7.38 Figure 7.39 Figure 7.40 Figure 7.41 Figure 7.42 Figure 7.43

PM10 Yearly Average, 1997 PM10 Monthly Average, 1997 SO2 Yearly Average, 1997 PM10, NO2 and SO2 Yearly Averages, 1997 O3 Yearly Average, 1997 O3 Monthly Average, 1997 NO Yearly Average, 1997 NO2 Yearly Average, 1997 NOx and Ozone Trends, 1997 NOx and Ozone Trends (excluding Bury), 1997 Temperature Yearly Average, 1997 Relative Humidity Yearly Average, 1997 Wind Speed Yearly Average, 1997 Daily Rainfall, 1997 Monthly Rainfall, 1997 PM10 Diurnal Average, Winter 1997 PM10 Diurnal Average, Summer 1997 SO2 Diurnal Average, Winter 1997 SO2 Diurnal Average, Summer 1997 Ozone Diurnal Variation, Winter 1997 Ozone Diurnal Variation, Summer 1997 NO Diurnal Variation, Winter 1997 NO Diurnal Variation, Summer 1997 NO2 Diurnal Variation, Winter 1997 NO2 Diurnal Variation, Summer 1997 Temperature Diurnal Average, Winter 1997 Temperature Diurnal Average, Summer 1997 Statistical Analyses and Durbin Watson Test Results PM10 vs. SO2, 1997 PM10 vs. NO2, 1997 PM10 vs. Temperature, 1997 PM10 vs. Relative Humidity, 1997 PM10 vs. Wind Speed, 1997 Ozone vs. NO, 1997 Ozone vs. NO2, 1997 Ozone vs. Temperature, 1997 Ozone vs. Relative Humidity, 1997 Ozone vs. Wind Speed, 1997 Prevailing Wind Direction, 1997 PM10 vs. Wind Direction, 1997 SO2 vs. Wind Direction, 1997 NO2 vs. Wind Direction, 1997 O3 vs. Wind Direction, 1997

156 157 158 158 159 160 161 161 162 163 163 164 164 165 165 167 167 168 168 169 169 170 170 171 171 172 172 173 174 174 175 175 176 176 177 177 178 178 179 180 180 181 181

15

Figure 7.44 Table 8.1 Figure 9.1 Figure 10.2

NO vs. Wind Direction, 1997 Chemical Composition of PM10 in the GMA Temperature Average in the MCMA and the GMA Relative Humidity and Rain Averages for the MCMA and the GMA Figure 9.3 Wind Speed in the MCMA and the GMA Figure 9.4 PM10 Average in the MCMA and the GMA Table 9.6 Chemical Composition of PM2.5 in the MCMA and the GMA Figure 9.5 O3 Average in the MCMA and the GMA

182 191 202 203 205 206 207 209

16

Abstract
Time series of particulate matter under ten micrometers (PM10) and ozone (O3) in the Mexico City Metropolitan Area (MCMA) and the Greater Manchester Area (GMA) were analysed in terms of the association with meteorological parameters and the time series of their precursor species. A description of the urban areas and the factors affecting the concentration of pollutants within their different atmospheres was prepared, while the data from the monitoring networks measuring the relevant pollutants in both areas was examined. In the MCMA, the results of the analyses suggested that PM10 were affected by temperature, relative humidity, pluvial precipitation and, under the higher regime, by wind speed. The majority of the PM10 were primary, and the small percentage of secondary particles was mainly associated with the oxidation of NO2. Ozone concentrations were also affected by the same meteorological parameters. Its concentration was controlled by the photolytic reaction of NO 2 , while it was predominantly removed through wet chemistry. In the GMA, relative humidity and pluvial precipitation did not show any influence in the concentration of either PM10 or O3. The rest of the parameters, however, had a strong influence on the loadings of both pollutants. Particles in the GMA were mostly secondary and formed through the oxidation of NO2. Ozone concentrations were primarily controlled by its reaction with NO.

17

Introduction
The main purpose of this thesis is to establish the formation and transformation mechanisms of particulate matter under ten micrometers (PM10) and ozone (O3) in the Mexico City Metropolitan Area (MCMA) and the Greater Manchester Area (GMA). This will be done through a thorough examination of the time series of both pollutants, their precursor species and of meteorological parameters. The different associations between all these and their chemistry and physics will be used to explain their seasonal distributions, concentration loadings and the interaction between the particles, ozone and the atmosphere of both places. In the first chapter, the boundary layer will be explored with the purpose of understanding the different forces affecting the atmosphere. In chapter two, the formation and transformation of aerosols will be reviewed, while chapter three will concentrate on the photochemical reactions affecting the tropospheric concentration of ozone.The two metropolitan areas will be described in chapter four, emphasising the factors influencing the levels of pollution in both urban agglomerations. The monitoring networks and the methods used to measure the different pollutants and meteorological parameters will also be reviewed in the same chapter. All the results of the analysis performed in this work will be presented in chapter five for the MCMA and chapter seven for the GMA. Chapter six will discuss the results of the analysis pertinent to the MCMA, while the same will be done for the GMA in chapter eight. A comparison of the main results of both areas will be undertaken in chapter nine. Finally, all the important conclusions drawn throughout the thesis will be included in the last chapter.

19

Boundary Layer Meteorology


Chapter One

1 Introduction
The importance of certain meteorological parameters on the way pollutants disperse is paramount. Wind speed and direction coupled with turbulence determine not only their vertical and horizontal movement, but also the velocity at which they disperse. In this chapter, we discuss the manner in which these parameters evolve and affect the lower part of the troposphere, the boundary layer. The input of energy from the sun and the way it is distributed throughout the globe is presented in the first section. The relationship between temperature, pressure and their respective gradients is then discussed, along with the effect of these factors on the forcings driving the motions within boundary layer. Wind is examined with the purpose of understanding the manner in which it develops and the way it distributes the incoming solar energy. Atmospheric pressure is presented with the aim of comprehending the thermal properties that it induces and its predominance on wind direction. Atmospheric stability is inherently linked to the buoyancy forces instigated by both the thermal and pressure properties of the earth. Section 1.32 presents the information related to this association and investigates the effect that it has on the stratification and depth of the different layers within the lower troposphere. Overall, all the information presented in this chapter aims to provide a framework for the investigation of the influence that the boundary layer, the atmosphere, and meteorology in general, have on the formation and transformation mechanisms of PM10 and Ozone in the metropolitan environments of the Mexico City Metropolitan Area and Greater Manchester Area.

1.1 Radiation
Energy from the sun is in the form of electromagnetic radiation; that is, energy derived from oscillating magnetic and electrostatic fields (Lyons and Scott, 1990). This type of radiation is capable of being transmitted through space and thus travels to reach and warm the earth, providing its only source of energy. The overall temperature of the earth remains fairly constant, suggesting that the earth and its atmosphere as a whole lose as much energy by radiation back into space as they receive from the sun. The amount of energy radiated from a body depends largely on the temperature of the body. Empirically, it has been shown that at a given temperature, there is a maximum amount of radiant energy that can be emitted per unit area of the body. This maximum is called the black body radiation. Similarly, a body that radiates at every wavelength the maximum possible intensity of radiation at a certain temperature is termed a black body. The sun radiates over the entire electromagnetic spectrum, although most of the energy is concentrated near the visible portion of the spectrum, the band of wavelengths from 400 to 700 nm.
23

The sun, although not a true black body, approximates a black body at a temperature of 6000 K, with a maximum intensity of incident radiation occurring in the visible region of the spectrum at 480 nm. The earth, on the other hand, resembles a black body at a temperature of 300 K, with a peak intensity occurring in the infrared region of the spectrum, at 10,000 nm. The latter quantities, the maximum intensity wavelengths, are obtained by combining several of the laws of radiation. The wavelength at which the maximum amount of radiation is emitted by a black body is found by substituting the temperature and solving for . (1. 1)

Where max is the maximum intensity wavelength and T is the temperature in degrees K. Solving equation 1.1, the black body temperature for the sun and the earth, T equal to 6000 and 300 K respectively, yields approximately 480 nm for the sun and 10,000 nm for the earths maximum emitting intensity wavelength. Kirchhoffs Law states that the absorptivity of a surface is equal to the emissivity for the same frequency and temperature. That is, a good absorber is also a good emitter and: (1.2 ) (Seinfeld, 1986) The absorptive properties of gases in the atmosphere are one of the most important aspects of both global meteorology and atmospheric chemistry. The most significant absorbing species are oxygen (O2), ozone (O3), water vapour, carbon dioxide (CO2), and dust. The absorption spectra of these gases are very complicated, but they do indicate the regions where the atmosphere is relatively transparent or opaque to radiation, where it can be transmitted without an appreciable loss of energy. Absorption may be so strong in some spectral regions, that no solar energy reaches the surface of the earth, i.e. absorption by O2 and specially O3 is responsible for intercepting practically all the incident radiation with wavelengths shorter than 290 nm. By contrast, the atmosphere is relatively transparent in the range from 300 to 800 nm, forming an atmospheric window in the spectrum. About 10% of the solar energy is concentrated in the 400 to 700 nm band, allowing most of the incident short-wave radiation from the sun to penetrate and reach the surface. From 300 to 800 nm the atmosphere is essentially transparent to incoming short-wave radiation and from 800 to 2000 nm, the terrestrial long-wave radiation is moderately absorbed by water vapour and other molecules in the atmosphere.

24

Water vapour absorbs in a complicated manner and mostly in the region where the suns and the earths radiation overlap. It is important to note that the molecules responsible for the most pronounced absorption of both solar and terrestrial radiation are minor constituents in the atmosphere, i.e. not nitrogen or oxygen. Thus, ozone in the upper atmosphere effectively absorbs solar radiation below 290 nm, whereas water vapour and CO2 absorb much of the long-wave terrestrial radiation. Following Kirchhoffs law, it would follow that they would also emit large amounts of long-wave radiation back into the surface or into space.

1.11 Radiation Balance


The solar beam is the sole input of energy into the earth and its atmospheric system. The magnitude of this input, known as the solar constant, is 1367 Wm2 and when it is averaged over the top of the atmosphere on an annual basis, the spatial input is equal to 342 Wm-2. As radiation from the sun passes through the upper atmosphere, about 3% is absorbed by stratospheric ozone. On average, gases, particles and clouds absorb around 20% of the incoming radiation. The average reflectivity (albedo) of the earth and atmosphere is in the order of 30%-50%, usually taken to be 34%, with the clouds being responsible for most of this reflectivity. Most qualitative analysis, with the purpose of simplifying the radiative equilibrium, break down the earth-atmosphere system into three sub-systems: earth, atmosphere and space. These are used to separate the radiation budget in such a manner that the net radiation balances; that is, the input equals the output when all the sub-systems are taken into account. On the basis of 100 units of incoming solar radiation, 47 are absorbed by the earth, 34 are radiated back to space (25 by clouds, 7 units are scattered by the atmosphere and 2 reflected by the earth) and 19 units are absorbed by the atmosphere (see Table 1.1). Table 1.1 Energy Received from the Sun, 100 Units Units 47 34 19 Total:100 Source Sun Sun Sun Destination Absorbed by the earth Radiated back to space Absorbed by atmosphere

The earth radiates in the infrared portion of the electromagnetic spectrum (as its temperature is in the order of 285 300 K). On the basis of the same units, the earth gains 47 units from the sun, loses 33 by evaporation and convection to the atmosphere, loses another 119 as long-wave radiation (113

25

absorbed by the atmosphere and 6 to space) and gains 105 units from reradiation back to the earth from the atmosphere (47-33-119+105 = 0, Table 1.2). Table 1.2 Gains and Loses by the Earth Units 47 -33 -119 105 Total: 0 Source Sun Evaporation and Condensation Earth Atmosphere Destination Earth Atmosphere Atmosphere and Space Earth

The atmosphere gains 19 units from the short-wave radiation emitted by the sun, 113 from the long-wave radiation of the earth and 33 by evaporation and convection from the earth. It loses 105 units re-radiating back to the earth and 60 units back to space (19+113+33-105-60 = 0, Table 1.3). Table 1.3 Gains and Loses of Energy by the Atmosphere Units 19 113 -105 33 -60 Total: 0 Source Destination Sun Atmosphere Earth Atmosphere Atmosphere Earth Evaporation and Condensation Atmosphere Atmosphere Space (all quantities taken from Seinfeld, 1986)

The whole earth-atmosphere system is in radiative equilibrium because the solar input (100 units) is matched exactly by the sum of the short-wave scattering and reflection, and the long-wave emission from the earth and the atmosphere. The atmosphere, by absorbing such a high fraction of the longwave radiation of the earth, acts as an insulator keeping the heat near the surface of the earth, an effect known as the green house effect. Water, in both vapour and droplet form, is the principal agent of this effect, absorbing in both high and low energy portions of the infrared spectrum. Carbon dioxide is second in importance as it has a lower atmospheric concentration and its absorption in the infrared portion of the spectrum is in a narrow band near 15000 nm.

1.2 Temperature and Pressure


The layers of the atmosphere are classified in a number of different ways: according to their chemical properties, their density, their temperature, etc.

26

From an air pollution stand point, however, it is useful to classify them according to their thermal structure. Starting from the outer most layer they can be identified as: 1) Thermosphere. The uppermost layer. Molecular densities are of the order of 1013 molecules cm-3, as compared with 2.5 x 1019 at sea level. Intense ultraviolet radiation dissociates N2 and O2. Temperatures exceed 1000 K. 2) Mesosphere. Extends from 50 to 85 km above the surface, overwhich temperature decreases with altitude until it reaches 175 K, the coldest point in the atmosphere. 3) Stratosphere. Extends from the tropopause to approximately 50 km in altitude. Temperature is constant in the lower stratosphere and then increases with altitude owing to the absorption of short-wave radiation by ozone. At the stratopause (the top of the stratosphere), the temperature reaches 270 K. There is little vertical mixing in the stratosphere. 4) Troposphere. The layer closest to the ground extending to an altitude of 15 km over the equator and 10 km over the poles. Temperature decreases with height at a rate of about 6.5 C km-1. Vertical convection keeps the air relatively well mixed. (Seinfeld, 1986) The behaviour of the atmosphere, especially in the troposphere, is affected greatly by the pressure and temperature gradients that exist within this layer. It is the thermal profile of the lower troposphere that determines the vertical mixing or convection, while the pressure difference between two locations drives the advective forces that produce different weather patterns around the globe. The relationship between these two parameters is extremely important. The pressure at any height z is due to the weight of the air above, thus a change in pressure in the vertical direction, obeys the hydrostatic equation:

(1.3) (Seinfeld, 1986) where is the density of air and g is the acceleration due to gravity. In order to study the relationship between temperature and pressure in the lower atmosphere, or more specifically in the lower troposphere, hypothetical air parcels (masses of air that may deform as they move vertically in the atmosphere) are used. The concept of an air parcel is useful as long as the parcel is of such a critical size that the exchange of air molecules across its boundaries is small in relation to the total number of air molecules in the parcel. As such parcel rises, it expands to account for the lowering pressure, but it does so in such a manner that the exchange of heat between the parcel and the surrounding air is negligible, that is adiabatically. As the parcel expands, its temperature decreases and a temperature differential (not of pressure) is established. The reverse is also true, as an air parcel sinks in the atmosphere it cools by compression, responding to the higher pressures.

27

As will be explained later (section 1.32), this temperature difference is very important because it drives the buoyancy forces determining the turbulence within the atmospheric boundary layer. The relation of temperature of the parcel to that of the air determines whether the parcel will rise, fall or whether it will reach a point of equilibrium. The variation of temperature with height for a rising parcel of dry air that cools adiabatically, that is, with no exchange of heat with its surroundings, is determined by dry adiabatic lapse rate ( ):

(1.4) where is the heat capacity at constant pressure per unit mass of air. The quantity g / is a constant for a dry air equal to 1 C / 102.39 m or 0.976 C / 100 m. If the air contains water vapour, the heat capacity at constant pressure ( ) must be corrected by incorporating the ratio of the mass of water vapour to the mass of dry air. The resulting rate of decrease will be smaller, i.e. if 3% of the atmospheric pressure is due to water vapour, -dT / dz = 1C / 103 m. As the parcel continues to rise, it will cool until the partial pressure of water vapour equals the saturation vapour pressure of water, and condensation may occur. At this point, the process is no longer adiabatic because latent heat will be released and an exchange of heat will occur. The rate of cooling after this point, with moist air, will be less than for the dry air parcel. Since the saturation vapour pressure of water increases very rapidly with temperature, the lapse rate for the moist air is strongly dependent on temperature. In the tropics, with warm air, the wet lapse rate may be up to one third of , whereas in the cold polar regions there is little difference between the two rates. Since the atmosphere is seldom adiabatic, the potential temperature ( ) is used because it provides a better measure of the temperature of a parcel of air when it is brought adiabatically from one height to another. In other words, it is the temperature that a parcel of dry air would have if it was brought adiabatically from one height to another.

(1.5) (Seinfeld, 1986) where = / v, and v is the heat capacity at constant volume per unit of mass of air, P is the air pressure (usually taken to be the surface pressure, 100 kPa or 1000 mb), P0 is a reference pressure, and -( -1) / is equal to 0.286 (obtained by setting R = C P C V = 287 J Kg -1 K -1 and C P = 1004 J Kg-1 K-1 and then substituting into the parameter).

28

The gradient of with z may be expressed in the terms of the absolute . Since the magnitude of is relatively close to T: temperature T, and

or (1.6) (Seinfeld, 1986) This parameter becomes very useful because adiabatic temperature profiles, based on potential temperature, are vertical on Cartesian co-ordinates against height z, thereby facilitating such comparisons. In reality, the atmosphere is seldom adiabatic. Other processes such as solar heating and the wind may exert a stronger influence on the prevailing temperature profile than the falling and rising of air parcels. The gradient expressed in equation 1.7 may be used as a measure of the departure of the actual temperature profile from adiabatic conditions.

1.3 Boundary Layer Forcings


Although the total input and output of radiant energy to and from the earth are essentially in balance, they are not evenly distributed. The amount of energy reaching the surface depends, in part, on the nature of the surface, the degree of cloudiness, and the latitude. In the polar regions, where lower solar angles than in the tropics are found, the same amount of solar radiation must pass through a thicker atmosphere and intercept a larger surface. The heating of the surface and of the atmosphere is therefore uneven throughout the globe. This disproportionate distribution of energy leads to the large-scale air motions of the earth, in particular, transporting energy from the tropics towards the polar regions. In general, the tendency is for warm tropical air to rise and cold polar air to sink, with pole-ward and equatorial flows respectively.

1.31 Wind
Air in motion or wind, is the agent of the atmosphere redistributing radiant energy throughout the globe. The air is set in motion by a series of forces: The pressure gradient force (PGF), tending to impel air motion from areas of high to areas of low pressure. Friction, acting opposite to the wind direction and proportional, roughly, to the square of the wind speed. The Coriolis force (CF), caused by the rotation of the earth. In terms of air motion, the atmosphere can be divided vertically into two sections. The first is the upper boundary layer, which is partly defined by the fact that the CF completely balances the PGF. It is at this level that wind achieves geostrophic speed.

29

The second laye extends from the surface to an average height of 600 m, varying from about 500 to 600 m. Within this layer, the PGF is counter balanced by the frictional drag exerted by the surface of the earth (hence called the surface layer) and by the CF. (Seinfeld, 1986)

1.311 General Circulation of the Atmosphere


As the wind travels from areas of high pressure towards areas of low pressure, the Coriolis force deflects it to the right in the Northern Hemisphere and to the left in the Southern Hemisphere (when viewed from above). This produces a net direction, at geostrophic speeds, parallel to the isobars (lines of equal atmospheric pressure). On both sides of the equator, warm tropical air rises and moves poleward while cool northern (or southern) air flows towards the equator in the lower levels. The circulation does not extend all the way to the poles because radiative cooling of the upper poleward flow causes it to subside at about 30 North and South latitude. The Coriolis force acting on these cells (air travelling from north to south in the Northern Hemisphere and reversed in the Southern Hemisphere) leads to easterly winds, more commonly known as the trade winds. Meanwhile, not all the subsiding air moves equatorward, but some travels north and is deflected towards the west resulting in westerly winds, or the westerlies. A similar situation occurs near the poles where warm air from the temperate zones moves poleward in the upper levels, eventually cooling by radiation and subsiding at the poles and resulting in the polar easterlies. At the boundaries of the subsiding zones, 0, 30 and 60 N and S latitude, the air is calm and little large-scale movement of air is generally registered. In the upper boundary layer, where the flow is only influenced by the PGF and the CF, the geostrophic wind speed (u g ) can be calculated by:

(Hare, 1989) where is the rate of rotation of the earth (7.3 x 10-5 rad s-1), is the the density of air, ug is the wind velocity (parallel to isobars latitude, (m s-1)) and dp / dn is the pressure gradient. The general pattern of circulation described above is not a completely realistic representation of the actual atmospheric circulation in any given day. The Coriolis force, along with the normal irregularities of the surface layer and its temperature, tend to disrupt the aforementioned system. In particular the CF and the turning of the wind vector towards regions of lower atmospheric pressure (explained in the next section) act on low level winds producing a vortex like circular motion. Air converging at low-pressure atmospheric systems (cyclones) will produce an inward spiralling motion, moving counter-clockwise in the Northern Hemisphere and clockwise in the Southern Hemisphere.

30

Similarly, air diverging from high-pressure atmospheric systems (anticyclones) will move outwards from the system, rotating clockwise in the Northern Hemisphere and counter-clockwise in the Southern Hemisphere. Typically, the centre of a cyclone will be a rising column of warm air and for an anticyclone, the reverse, a sinking column of cold air. In general, cyclones are associated with cloudy skies and severe weather, many times causing precipitation. In contrast, anticyclones are associated with clear skies, light winds and fair weather often leading to severe thermal inversions. These systems become important as they drive part of the global climate. Some are semi-permanent as they shift position only slightly during an annual cycle, while the majority are migrant and temporary. In particular, the semi-permanent subtropical anticyclones centred over the major oceans produce warm humid climate on the eastern side of continents and dry desert like ecosystems on the western side. Good examples include the desert of the Sahara, Western Australia and Southern California on the west side, and the rainforest in South America and Africa for the eastern side.

1.312 Surface Wind


Wind in the surface layer is largely controlled by the frictional drag imposed on the flow by the underlying surface. This stress retards motion close to the ground and gives rise to a sharp decrease of mean horizontal wind speed as the surface is approached. To a large extent, the depth of this frictional influence depends on the roughness of the surface. The depth increases with greater roughness; hence, the vertical gradient of mean wind speed is greatest over smooth terrain, and least over rough surfaces. In light winds, the depth also depends upon the amount of thermal convection generated at the surface. With strong surface heating, the height is greater, while during surface cooling it is less. The surface layer of the atmosphere is also defined by the balance of the three forces described above. As in the upper boundary layer, the CF balances the PGF but here, the frictional drag or shear stress shifts the balance, reducing the influence of the CF. In terms of vectors, the PGF must be directed from high to low atmospheric pressure, the frictional force must be opposite to the wind velocity (u), and a balance can be achieved only if the wind is directed at some angle towards the region of low pressure (instead of parallel to the isobar, as with the geostrophic wind in the upper boundary layer). Figures 1.1a and 1.1b illustrate the difference between the two levels of influence. The angle of the surface wind u (measured at 10 m above the surface) to the isobars is of the order of 45 in build up areas, and the speed is approximately fifty percent of the geostrophic wind. This angle increases and the wind speed decreases as the ground is approached since the frictional force or shear stress increases.

31

Figure 1.1a and 1.1b Variation of wind direction with altitude. (a) Balance in the upper atmosphere, with geostrophic wind CF = PGF (b) Balance in the surface layer CF +F = PGF

Low Pressure

(a)

(b) u = Wind Vector

u = Wind Vector F
Coriolis

friction

Coriolis

High Pressure
Vectors not to scale

(Adapted from Seinfeld, 1986 pg.) The actual profile of the wind gradient with height under neutral stability (i.e. atmospheric stability classification in which buoyancy is neither enhanced nor opposed) is described by a logarithmic profile, following the relationship:

(1.9) Where is the mean wind speed (m s-1) at height z, u* is the frictional velocity (m s-1), k the von Karmans constant ( 0.40) and z0 the roughness length (m). The length z0 is a measure of the aerodynamic roughness of the surface. Typical values are given in Table 1.4. The value of u* is usually obtained by measuring the wind velocity at a reference height z (commonly 10 m) and then substituting the values and solving equation 1.9. Table 1.4 Roughness Lengths for various surfaces Surface z0 (m)

Very smooth (ice, mud flats) 10-5 Snow 10-3 Smooth sea 10-3 Level desert 10-3 Lawn 10-2 Uncut grass 0.05 Fully grown root crops 0.1 Tree covered 1 Low-density residential 2 Central business district 5-10 (McRae et al., 1982; cited in Seinfeld, 1986, pg. 495)

32

1.32 Atmospheric Stability


Temperature profiles in the lower atmosphere determine in part the stability of the atmosphere, or in other words, the degree to which wind induced turbulence, surface roughness, or buoyancy will propagate through the layer. Roughly, the atmosphere can be classified into three main categories: neutral, stable and unstable. Under neutral stability the environmental lapse rate ( ) is equal to the dry adiabatic lapse rate ( ) and buoyancy is neither enhanced nor opposed. Under strongly stable conditions, disturbances are highly damped and mixing of species is strongly suppressed. By contrast, unstable conditions lead to a buoyant atmosphere and mixing is enhanced. There are many ways of defining stability in the atmosphere but they all refer, at some stage, to the environmental and the dry adiabatic lapse rates , and respectively. Seinfeld (1986) uses what is probably one of the simplest, defining stability according to buoyancy forces and acceleration due to gravity. He defines the acceleration experienced by a parcel of air as follows:

for To To Where T denotes the temperature of the air parcel, T the temperature of the surrounding air and g is acceleration due to gravity. Substituting for dT/dz and for dT/dz equation 1.10 can be redefined as:

(1.11) Under these conditions, a parcel of air will rise and continue to rise as long as > . Similarly, a parcel of air cooler than the surrounding air will descend and continue to descend if its rate of heating is less than the lapse rate in the atmosphere ( < ). In summary, the stability of the atmosphere is defined as follows: = Neutral Stability > Unstable (vertical motions enhanced) < Stable (vertical motions suppressed) There are many other ways to define stability but as mentioned, this is one of the simplest and most useful. The actual environmental lapse rate can, in most cases, be readily measured and the dry adiabatic lapse rate is constant at 9.8 K for every 1000 m (roughly 1 C for every 100 m of rise). Lyons and Scott (1990) define stability in terms of the parameter S, defined as:

33

(1.12) Where T is the environmental temperature and is the potential temperature (equation 1.5). Following this equation, the air is said to be stable when S is greater than zero, meaning that the buoyancy force is downward and vertical motions are suppressed. When S is less than zero there is a positive buoyancy force and the atmosphere is said to be unstable. In cases where S is equal to zero, the atmosphere is classified as neutral or simply dry adiabatic.

1.33 Boundary Layer Depth and Structure


Drawing on the information presented before, the boundary layer (BL) can now be redefined as the part of the troposphere that is directly influenced by the presence of the earths surface, and responds to surface forcings with a time-scale of about an hour or less (Stull, 1988). As discussed in section 1.311, the atmosphere is controlled, in part, by pressure systems and the manner in which these help to establish the general circulation of air. Within the BL, these systems influence the height of the layer by either drawing air from aloft (subsiding into anticyclone systems) or by converging air from the surface (updrafts away from cyclone systems). Over both land and oceans, the general nature of the BL is to be thinner under high-pressure systems than in low-pressure regions. The subsiding air in anticyclone systems creates diverging horizontal air motions at the surface, which in turn result in a downward movement of the BL. Opposite to this effect, in low-pressure areas the upward movement of air from the surface carries the BL away from the ground into large altitudes. Over land, in high-pressure regions, the BL has a well-defined structure that evolves with the diurnal cycle. In the early morning, the air tends to be stable but as the sun begins to heat the ground the environmental lapse rate increases rapidly, until a shallow layer with a lapse rate greater than the dry adiabatic is created. This turbulent layer progresses throughout the day and may become one or two km deep. If the air is moist, clouds may form on many of the ascending thermals and these may rise to considerable heights, especially if the upper troposphere is cool. As evening progresses, however, the solar heating at the ground level ceases and conditions become more stable again. Vertical motions in the atmosphere are generally limited by thermal inversions, regions of very stable air in which the temperature actually increases with height. Inversions correspond to a negative lapse rate and are the opposite of lapse conditions in which the normal, negative temperature gradient prevails. Usually, there is little wind, and pollutants emitted into the inverted layer are slow to disperse.

34

Stable conditions are common during the night, and in winter may persist all day leading to severe thermal inversions. Based on the diurnal cycle described above, the BL can be further characterised as having a mixed layer (ML), a residual layer (RL), and a stable layer (SL). When clouds are present, the ML is subdivided into a cloud layer (CL) and a sub-cloud layer (SCL). Under this scheme, the surface layer is the lowest portion of the BL, where shear turbulent fluxes and shear stress vary by less than 10% of their magnitude, that is they remain fairly constant (Stull, 1988).

1.331 Mixed Layer


Turbulence in the mixed layer is usually convectively driven and as such, it is also termed the convective layer. It is possible, however, to have a well mixed layer in regions of strong wind where advection is also important. The convective forcings mainly include heat transfer from a sun-heated surface and radiative cooling from the top of the cloud base layer (when present). The former situation creates thermals of warm air rising through the layer while the latter, cool subsiding air from the clouds. Both situations may coexist, especially when a cool cloud topped ML is being advected over a warmer surface. On basically cloud free days, the ML growth is bound to solar heating. Half an hour after sunrise, the ML begins to grow and is usually characterised by intense mixing in a statically unstable situation where thermals of warm air rise from the ground. It reaches its maximum height in the late afternoon, growing by entrainment into the less turbulent air from above. Temperature profiles in the middle portion of the ML tend to be adiabatic while near the surface they tend to be super-adiabatic. A stable layer (frequently called an inversion layer) limits the ML acting as a cap or lid to rising thermals. The boundary between the ML and the SL is called the entrainment zone, as it occurs there. Most pollutants are emitted near the surface and transported upward by eddies, such as thermals, throughout the ML. The inability of thermals to penetrate deep into the SL results in a trapping of pollutants below the inversion layer. This situation often occurs under high-pressure systems, and produces severe pollution episodes (Stull, 1988). On overcast days insolation is reduced at the ground, causing a reduction of the intensity of the thermals and lower ML growth. If the clouds are thick enough, the ML may even become non-turbulent or neutrally stratified.

1.332 Residual Layer


In the afternoon, approximately half an hour before sunset, thermals induced by surface heating stop forming and turbulence begins to decay. The new forming atmospheric layer may still have some of the attributes of the ML and as such, it is termed the Residual Layer (RL). 35

In general, when it is well established, the RL is neutrally stable, as turbulence is approximately equal in all directions. In terms of temperature profiles, the rate of decrease with height is slow and close to adiabatic values. This layer can be understood of as a transition between the ML and the Stable Layer.

1.333 Stable Layer


As the night progresses, the bottom portion of the RL is transformed by its contact with the ground into a stable layer, the Stable Boundary Layer (SL). This layer is characterised by sporadic turbulence and statically stable air. Wind regimens near the ground tend to be light, but aloft they may achieve super-geostrophic speeds in a phenomenon called the low-level jet or the nocturnal jet. The combination of the nocturnal jet and the statically stable air near the ground results in short bursts of turbulence that in some cases may cause mixing within the SL. During non-turbulent times, the two levels become virtually decoupled. Wind patterns exhibit a very complex behaviour during the night. They tend to be light near the surface, but at altitudes on the order of 200 m, the nocturnal jet may reach speeds of 10 30 m / s. A few hundred metres above, wind speed decreases to approximately geostrophic values. In this layer, the lapse rate is positive and the temperature increases with height, causing thermal inversions that may trap pollutants near the ground. Although the SL is normally established during the night, it may also form during the day, as long as the underlying surface is colder than the air, as it often occurs when warm air is advected over a colder surface. Unlike the ML, the SL has no clearly defined upper boundary and its limit is poorly defined because it blends smoothly into the RL above. Figures 1.2 and 1.3 show the diurnal evolution of the layers just described. The first is a diagram of the different layers of the boundary layer in a high-pressure system over land. The second illustrates the virtual potential temperature profiles, an analogous parameter to potential temperature, and indicates the progression of the layers. Figure 1.3 was generated with rawinsonde soundings at the time S indicated in Figure 1.2. Figure 1.2 Boundary Layer in a high-pressure system over land
2000 Entrainment Zone Free Atmosphere Capping Inversion Entrainment Zone 1000 Mixed Layer Stable Layer Noon Suns Midnight Sunrise Noon Mixed Layer Mixed Layer

0 S1 S2

S3

S4

S5

S6

36

Figure 1.3 Profiles of mean virtual potential temperature, Ov

S1 Z FA Z

S2 Z FA

S3 FA

RL ML V S4 Z FA RL SL ML ML V V Z FA RL S5 Z SL V S6

RL

SL V

FA CL

ML

SL

This figure shows the boundary-Layer evolution during a diurnal cycle starting at about 1600 local time. S1- -S6 identifies each sounding with an associated launch time indicated in Figure 1.2. (Adapted from Stull, 1988 pg. 17)

1.4 Urban Boundary Layer


Large urban agglomerations are anthropogenic sources of heat and pollution generated, to a large extent, by the burning of fossil fuels. These areas are also distinguished by altered environments with rough surfaces and materials that have high reflectivities and heat capacities. In general, the processes of urbanisation involve the transformation of the radiative, thermal, moisture and aerodynamic characteristics of the natural environment, thereby dislocating the radiative balance and the atmospheric properties of the region (Oke, 1987). As air flows from a rural to an urban environment it encounters completely different boundary conditions and as such its BL is identified independently as an internal boundary layer or the Urban Boundary Layer (UBL). The properties of this layer are dependent upon the individual characteristics of each city but in general, they are governed by their size, location and meteorological conditions. According to Stull (1988), the UBL can be subdivided into four strata. The portion of the BL between the rooftops and the ground is known as the urban canopy layer, analogous to the plant canopy in a forest. Within this region are urban canyons or streets, ducting and trapping airflow throughout the city. Above this is the turbulent wake layer, where the individual wakes and the internal fluxes of momentum in between the buildings and surface patterns can still be distinguished. 37

The next stratum, the surface layer, is where individual fluxes and wakes are not important but their average effects are still discernable. Finally, the urban mixed layer extends to the top of the UBL, which itself might be an internal BL within the large-scale flows. Figure 1.4 shows the different layers of the UBL. Figure 1.4 Urban Boundary Layer Structure
Z
1

Wind

Mixed Layer Surface Layer

BL

Wake Layer Canopy Layers within the urban boundary layer, (adapted from Stull, 1988, pg 612)

(Adapted from Stull, 1998 pg.11)

1.41 Urban Heat Island


Incoming solar radiation in the UBL is usually reduced by the blanket of pollutants found within cities. The level of attenuation varies considerably and is dependent upon, amongst other things, the type of pollutant and its atmospheric concentration. In heavily industrialised urban settlements where coal burning is dominant, a 10% to 20% reduction is normal. On the other hand, in less industrialised areas, where vehicle pollution is important and photochemical pollution is dominant, 2% 10% attenuation can be expected. These ranges are linked to the seasonal variations of pollutant concentrations and may, on heavily polluted days, exceed 30% attenuation (Oke, 1987). Generally, the deficit in solar radiation is compensated by lower albedos (usually in the order of 0.15) than the surrounding countryside and an increased absorption of the radiation by surface materials. Within cities the ground is generally covered with a large percentage of asphalt and concrete, which are usually dry, waterproof surfaces with albedos and heat capacities that convert and store incoming solar radiation as sensible heat better than the surrounding rural areas. The net result of the increased capacity to store and convert heat and the lower reflection rates is to increase the temperature of urban areas. This warming is known as the Heat Island Effect (HI). The form and size of the phenomenon varies in time and space as a result of meteorological, locational and urban characteristics.

38

The intensity of the HI (Tu-r), or the thermal difference between the rural to urban environments, is greatest under stable anticyclonic conditions, that is, under light or calm wind speeds and cloudless skies (Oke, 1987). These conditions are usually encountered in the nocturnal cycle of the day hence, the HI is best developed over night (see Figure 1.5), although it can also occur during the day. Figure 1.5 Diurnal Variation in the Urban and Rural boundary Layers
20

15

10

R ura l

U rba n

T (C)
5 0 12 6 0 6 12

Local Time (h)

(Adapted from Stull, 1988 pg. 610) At night, when the HI is well developed, Tu-r is inversely related to the wind speed (u) and cloud cover. As the wind speed intensifies, the thermal differences between the urban and rural boundaries dissipate and the horizontal gradients disappear. Based on observations, it appears that a critical point of u is in the order of 9 m s-1 for a city with a population of about one million and about 5 to 2 m s -1 for populations of 10 5 and 10 4 respectively (Oke, 1987). As it becomes obvious from these relations, the HI effect is also related to the size of the city. Using population (P) as a surrogate of city size, Tu-r is proportional to the Log of P, so the larger the population, or the city, the stronger the intensity of the heat island effect. The intensity of the HI effect is usually in the order of 2-5 C but Tu-r as large as 12 C have been recorded (Oke, 1987). Another effect common in urban areas is a reduction of the relative humidity. Although not enough data is available, it is speculated that it relates to the waterproofing of the surfaces and the replacement of green areas with cement or asphalt (reducing or even displacing evapotranspiration from plants).

39

1.42 Wind
As stated before, the surface roughness within cities is much higher than in their rural surroundings. Some of the features within urban areas (i.e. tall buildings with sharp edges) have roughness lengths that have been documented as the roughest of all elements. Table 1.5 compares some values of the roughness lengths in urban elements. Table 1.5 Typical roughness length (Z0) of urban terrain Terrain Scattered settlement (farms, villages, trees, hedges) Urban Suburban Low density residences and gardens High density High density, less than 5 storey row and block buildings. High density plus multi-storey blocks. (Z0) 0.2 0.6

0.4 1.2 0.8 1.8 1.5 2.5 2.5 - 10 (Oke, 1987)

The effect of the larger roughness length is to retard wind flow within the cities and to create more mechanically induced turbulence. Wind speed in the urban canopy is usually reduced. Two exceptions apply: Faster moving upper air layers are either deflected downwards by the relatively tall buildings or are channelled into jets along the streets oriented in the same direction as the flow. In a well developed HI, the horizontal temperature (and therefore of pressure) gradients across the urban / rural boundary can be sufficient to induce low level breeze from the countryside into the city, in the same manner as a sea breeze. The flow converges upon the city centre from all directions. As in the general circulation system, this must result in uplift over the city core and a counter flow from the city to the countryside aloft. If the canopy portion of the inflow is strong enough to overcome the frictional drag of the canyon walls, then the wind speed may be slightly greater than in the surrounding rural areas. (Oke, 1987) In general, however, the tendency is to have lower wind speeds in urban areas than in the surrounding countryside.

40

1.43 Urban Boundary Layer Characteristics


There are several properties that characterise the UBL, but some of the most important are: the excess temperature induced by the reduction in radiative cooling; the enhanced turbulence produced by the excess temperature and the increased roughness length; the reduction of relative humidity; and the retardation of wind speed. The compounded effect that all these have is to change the overall climate in such a manner that it produces a dome like structure over the city, creating a microclimate or an amalgamation of microclimates, depending on the size of the city (Seinfeld, 1988). When the dome is well defined it is usually mixed inside, with the stratification mentioned in section 1.4. Turbulence within this internal boundary layer is both convectively and mechanically driven, with thermals and roughness elements promoting it. Moreover, as the air converges in the city centre and spreads out into the rural areas, in the circulation pattern mentioned in the last section, the height of the UBL is pushed upwards, increasing the volume in which atmospheric pollutants disperse. The height of the UBL is in part determined by solar heating and as such grows during the day and contracts during the night, just as the ML in less populated areas. The layer just above the UBL tends to be statically stable, acting as a cap or lid to the UBL. Potential temperature profiles are shown in Figure 1.6. Figure 1. 6 Potential Temperature Profiles

Z
(a) Day
Urban Boundary Layer

Wind

Urban Plume

Z
(b) Night Wind
Urban Plume

Urban Boundary Layer and urban plume for a windy (a) day, and (b) night. (Adapted from Stull, 1998 pg 611)

(Adapted from Stull, 1988, pg. 611)

41

On windy days, the excess temperature, pollutants and deficits of humidity are carried downwind in an urban plume (Figure 1.6). These plumes may be as wide as the city itself and can be transported hundreds of kilometres downwind. On less windy days air circulation is only within the UBL and in general is much slower than in the adjacent rural areas.

1.5 Conclusion
The information presented in this chapter pertains to the extent of the effect that the atmosphere has on the concentration of air pollutants. The level of influence extends from the only source of energy available to the earth, to the way this energy affects the stability of the atmosphere and the depth of the Boundary Layer. Temperature and pressure have a profound effect as they drive the convective and advective forces that in turn govern the movement of air in the horizontal and vertical direction. Temperature profiles in the BL are established because of the pressure gradient and in turn, they propel the different stability classes in the atmosphere. In more than one way, these classes determine the possibility that pollutants have to disperse. Most air pollutants are emitted at the surface and their ability to disperse throughout the atmosphere is dependent on eddies, such as thermals, to transport them aloft and then across a wind field. Furthermore, the stability classes in conjunction with the different pressure systems of the atmosphere, namely cyclones and anticyclones, determine the height of the mixed layer and thereby the volume of air in which the pollutants will be diluted and dispersed. When the boundary layer is turbulent and deep, mixing will occur and the pollutants concentration will be less than in a shallow and stable layer. To conclude, it is the combination of atmospheric stability and wind speed and direction that determine, to some extent, the state of the atmosphere. This in turn affects the manner in which pollutants disperse and as a consequence their inherent concentration within the boundaries of space and time.

42

Atmospheric Aerosols
Chapter Two

2 Introduction
The concept of an atmospheric aerosol was first introduced by Schmauss and Wigand (1929) as an assembly of suspended particles in the air (Warneck, 1988). Atmospheric aerosols are defined as any substance, except for pure water, that exists as a liquid or solid in the atmosphere under normal conditions, and are of microscopic or sub-microscopic size but larger than molecular size (about 2 ). Among atmospheric constituents, aerosols are unique in their complexity: either they arise from direct emissions as primary particles or they condense into pre-existing material or they nucleate into new particles (secondary particulates, Seinfeld, 1986). In this chapter, the different aerosol classification schemes will be introduced along with some of the most important formation and transformation mechanisms. Some of the deleterious effects will be mentioned and the removal pathways will be highlighted.

2.1 Atmospheric Aerosols


Natural aerosols are mainly composed of material originating either from the crust of the earth (including its vegetation and other ecosystems) or from the oceans. Both have considerable sources and as such have their own classification, continental and maritime aerosols respectively. In pollution terms however, the categories had to be expanded to include the urban species especially as human activities are producing large amounts of pollutants that, through different mechanisms, may end up as suspended material. Urban aerosols are mixtures of primary emissions from industrial, transport and power generation activities, other natural sources and secondary material formed through chemical transformations in the atmosphere (Seinfeld and Pandis, 1998). Among the different properties of aerosols, size is one of the most important. It determines, in part, their effects and fate; from their formation and transformation processes, their residence time in the atmosphere to their light scattering potential and deposition patterns in the lungs. Size distributions of atmospheric particles are therefore a very important parameter. They help to characterise aerosols analytically and hence to aid in their analysis. The basic distributions are agglomerated in number, and mass. For most purposes, mass and volume distributions can be taken to be the same. When these two distributions are analysed together, they show that most particles are quite small, below 0.1 m, while most of the volume is found in particles larger than 0.1 m. Volume (and therefore mass) distributions tend to be bimodal, with a minimum around 1 and 3m. Particles larger than this are termed coarse, while the smaller are called fine. Willeke and Whitby (1975, cited in U.S. E.P.A, 1996a pg. 3-5) identified three modes: nuclei, accumulation and coarse. The smallest (nuclei) corresponds to particles below 0.1 m, the middle mode (accumulation) ranges between 0.1 and 2 m, while the larger (coarse) are those exceeding 2 m.

45

Fine particles include the two smaller modes while the coarse is the same in both characterisations. The boundaries between these modes are not precise, but they give an idea of particle size characteristics. Summarising and expanding on the above named distributions, aerosol scientists use three conventions to classify particles by size. These schemes are based on different properties of their size, monitoring equipment, health hazards and research interests (U.S.E.P.A, 1996a). They are represented as: Mode: Based on formation mechanisms and the modal distribution observed in the atmosphere, i.e. nuclei, accumulation and coarse modes. Cut Point: Based on the 50% cut point of the specific sampling devices, i.e.PM2.5, PM10 and PM10-2.5. Dosimetry: Based on the ability of the particles to entercertain regions of the respiratory system, i.e. inhalable, thoracic, respirable. (U.S. E.P.A., 1996a) In conjunction with the above named schemes, suspended particles are also categorised according to their aerodynamic size spectra. This size classification is greatly facilitated by assuming a spherical shape as they can then be defined by their aerodynamic radius or diameter. Two parameters that are often used are the Stokes diameter (Dp) and the aerodynamic diameter (Da). The Stokes diameter describes particle size based on the aerodynamic drag force imparted on a particle when its velocity differs from that of a surrounding fluid. For a smooth, spherically shaped particle, Dp is exactly the physical diameter of the particle. For irregular shaped particles, Dp is the diameter of an equivalent sphere that would have the same aerodynamic resistance. Aerodynamic diameter depends on particle density and is defined as the diameter of a spherical particle with equal settling velocity but a material density of 1 g/cm3. For particles greater than about 0.5 m, the aerodynamic diameter is generally the quantity of interest because it is the parameter important in particle transport, collection and respiratory tract deposition. Most sampling methods are based on particle aerodynamic diameter (U.S. E.P.A, 1996a). Aerodynamic diameter, Da, is related to Dp by: (2.1) Where is the particle density, and C and Ca are the Cunningham slip factors evaluated for particle parameters Dp and Da respectively (U.S. E.P.A, 1996a). The slip factor is a function of the ratio between particle diameter and mean free path of the suspending gas; it is given by the expression (Hinds, 1982 cited in U.S. E.P.A, 1996a, pg. 3-8):

(2.2)

46

Where is the mean free path of air (at normal conditions, temperature = 20 C, pressure = 1 atmosphere, = 0.066 m). For large particles (Dp > 5 m) C = 1; while for smaller particles C >1. For particles with diameters greater than , the aerodynamic diameter given by equation 2.1 is approximated by:

(2.3) This expression, which shows that aerodynamic diameter is directly proportional to the square root of the particle density, is often used for particles larger than or equal to 0.5 m. For particles with diameters much smaller than , the slip factor must be taken into account. In this case the aerodynamic diameter is proportional to the particle density [Da = ( )Dp for Dp << ] (U.S. E.P.A., 1996a). In pollution terms and for the purpose of this project the aerosols of concern are those with an aerodynamic diameter smaller than ten micrometers (PM10). These have several properties that affect the health of humans and the ecosystems exposed to them. Within this size range, the aerosols can be deposited in the tracheobronchial and alveolar regions of the human respiratory system and they tend to scatter light in the visible part of the spectrum (Villalobos-Pietrini et al., 1995). PM10 particles are mainly of primary origin, but include the smaller range that are formed through chemical reactions and are secondary in nature. Concomitant to the PM10 range, environmental scientists make another important distinction, particles under 2.5 m (fine particles). These particles are important in urban environments as many of the pollutants transforming into aerosols will fall within this size range (U.S. E.P.A., 1996a).

2.2 Formation of Aerosols


Aerosols can be formed and transformed in the atmosphere via several mechanisms, some of which are inevitable and natural while others area result of human activity (anthropogenic). Some of the natural sources include, but are not limited to, soil erosion, forest fires, volcanic eruptions, and marine sea breeze. In the urban environment however, soil erosion and the combustion of fossil fuels are the two major sources of particulate matter. Aerosols in urban environments are mainly formed by nucleation or coagulation. The former adds new particles, while the later increases their size and changes their size distribution.

2.21 Nucleation
Formation of new particles in the atmosphere is dependent on the vapour pressure of condensible species, that is, one at supersaturation. 47

The saturation ratio, S, is defined as relation of the partial pressure of species, p, to its saturation vapour pressure, p 0 , so that the: (2.4) (Seinfeld and Pandis, 1998) As stated, for nucleation to occur, the species vapour pressure must exceed the saturation (or equilibrium) vapour pressure. A ratio less than one is considered to be undersaturated, equal to one means saturation and greater than one is equivalent to supersaturation. Species may condense into a new particle (homogeneous nucleation) through gas to particle conversion (GTP) or onto the surface of a pre-existing particle (heterogeneous nucleation). These two formation mechanisms are usually referred to as nucleation and condensation, respectively. In both cases, there needs to be supersaturation of the condensible species, either in relation to the pure substance or in relation to a solution. The prevailing mechanism will depend on, among other parameters, the available surface area of preexisting particles. In urban environments, where the surface area is large enough to rapidly scavenge the newly formed condensible species, new particle formation is not important except close to the emission sources. Simplifying the formation of new particles, nucleation involves the spontaneous burst of a condensible material into a different physical phase, i.e. liquid or solid. This process occurs when molecular clusters grow beyond a critical size to become embryos. At saturation levels (S = 1), clusters gain and lose monomer molecules at the same rate. In supersaturation conditions (S > 1), the molecular clusters gain more monomers than they lose, hence they reach their critical size and change the cluster size. At this stage, phase change may occur (Seinfeld and Pandis, 1998). Nucleation of new particles happens at very specific atmospheric conditions. First and most important, there needs to be supersaturation. In the case of a single species, saturation ratios need to be relatively high. Water vapour, for example, requires saturation above 5 (S > 5). These values seldom occur in the atmosphere and therefore condensation tends to prevail as a general particle formation mechanism. Using the same example, condensation would require a much lower saturation value, in the order of 1.05 (S = 1.05, Warneck, 1988). There are however, cases where nucleation is competitive and effective. Bimolecular nucleation or conucleation of two species can occur even when both vapours are undersaturated. The requirement under these conditions is that the individual species are saturated in relation to a solution and not to the pure substance. A good example of this process is the formation of the sulphuric aerosol through GTP. The procedure occurs because sulphur dioxide in the presence of water vapour has a much lower equilibrium vapour pressure than on its own.

48

Sulphur dioxide is oxidised in the presence of hydroxyl (OH) and hydroperoxyl (HO2) radicals (along with several hydrocarbon groups) when they are irradiated with natural or artificial light. The following reaction system exemplifies this process. (2.5) (2.6) (2.7) (Fox, 1986) Sulphuric acid gas (H2SO4) in the presence of water vapour has an extremely low saturation vapour pressure in relation to the solution. Oxidation of relatively small amounts of sulphur dioxide can result in a gas phase concentration of sulphuric acid exceeding the equilibrium vapour pressure in the atmosphere, and the subsequent formation of the sulphuric acid aerosol (Fox, 1986). Recent estimates of the saturation vapour pressure at which H2SO4 will nucleate into an aerosol stand in the order of 4.66 x10-4 mbar at 300 K and 3.33 x 10-5 mbar at 296 K (Warneck, 1988). Sulphur species are one the most important gas to particle conversion classes, although nitrogen species can also be significant. The principal mechanism for gas-phase production of nitrates is the reaction of OH with NO2 to form HNO3.

(2.8) The reaction of OH with NO2 is approximately ten times faster than the reaction of OH with SO2 (Finlayson-Pitts and Pitts, 1986 cited in U.S.E.P.A., 1996a, pg. 3-65). Therefore, NO2 is preferentially converted to HNO3, and the conversion of SO2 to H2SO4 is delayed until much of the NO2 has reacted. By contrast, much higher concentrations of HNO3 (roughly nine orders of magnitude) are necessary to produce the same nucleation rate as with H2SO4 (Seinfeld, 1986). Many organic gases are readily oxidised to forms that have low volatility and as such condense into the aerosol phase. In general, straight-chain alkenes, alkynes and carbonyl compounds do not generate aerosols but the ability of the alkene compounds to produce particles increases when the carbon number is greater than six. Most of the particles formed through GTP are secondary in nature and their formation depends on the concentration of gaseous species such as SO2, NO2, OH and HO2. Moreover, their formation depends on the atmospheric conditions, including solar radiation and relative humidity, and their interaction with precursors and pre-existing particles in the atmosphere.

49

2.22 Coagulation
Aerosols in the nuclei mode undergo collisions with each other due to their Brownian motion: the constant bombardment of molecules of the surrounding fluid, i.e. air. It causes random movement of the particles that, in the absence of external forces, causes them to collide with each other and grow to form new larger particles. This process, called coagulation, causes the size distribution of the particles to change towards larger sizes. For particles with a radius much greater than the mean free path of air, (0.066 m for atmospheric conditions) and whose motion is hampered by friction, the probability of collision is controlled by diffusion. Under these conditions the collision frequency is calculated by considering the flow of particles of radius r 2 to a fixed particle with radius r 1 , expressed as: (2.9) Where K(r1,r2) is the coagulation function, r1 and r2 are the radii of the particles and D1 and D2 are the diffusion coefficients associated with the two particles (Warneck, 1988). As an example, Peter Warneck (1988) published in table 7-4 pg. 289, several diffusion coefficients along with other parameters for particles of different radii. He reports values for particles with a diameter of 0.5m and 5m as 1.276x10-11 and 1.190x10-12 m2s-1, respectively. The above equation is significant because at ambient humidity conditions particles will be sheathed with moisture, therefore increasing the sticking probability and making it close to unity. This will then result in the formation of a new, larger particle. For aerosols closer to, or smaller than, the value of , more complicated equations and correction factors apply as they are either controlled by the laws of gas kinetics or undergo different particle flow characteristics (Warneck, 1988). Coagulation, therefore, is an efficient formation process for accumulation mode particles (nuclei growing into accumulation mode). As previously stated, it changes the size distribution of the atmospheric particles into the larger modes. The formation mechanisms presented in the preceding sections represent the most important in the gaseous phase. There are, however, other relevant formation pathways.

2.23 Aqueous Phase Reaction Mechanisms


The aqueous-phase conversion of dissolved SO2 to sulphate is thought to be the most important chemical transformation in cloud water. Dissolution of SO2 in water results in the formation of three radical species: hydrated SO2 50

(SO2H2O), the bisulphite ion (HSO3-) and the sulphite ion (SO3), all of which are termed S(IV) (dissolved sulphur in oxidation state IV) when they are referred to as a single entity. At the pH range of environmental interest (pH = 2-7) most of the S(IV) is in the form of HSO3-, where as at low pH (pH < 2), all the S(IV) occurs as SO2H2O. At higher pHs the preferred state of S(IV) is SO3= (U.S.E.P.A., 1996a). Dissolved sulphur in oxidation state VI, S(VI), is the term used for the entity agglomerating H2SO4(aq), HSO4- and SO4. The most significant pathways for the transformation of S(IV) to S(VI) include oxidation of S(IV) via reactions with O3, H2O2, O2 (catalysed by Mn2+ and Fe3+) and NO2. Ozone reacts with SO2 in the gas phase very slowly, the aqueous reaction however, is fast. (2.10) (U.S.E.P.A., 1996a) This reaction is significant when the pH exceeds four. Hydrogen peroxide, H2O2, is one of the most effective oxidants of S(IV) in clouds and fogs, as it is very soluble in water and under normal ambient conditions its aqueous-phase concentration is approximately six orders of magnitude higher than that of ozone.

This reaction is very fast: field studies have suggested that H2O2 (g) and SO2 (g) rarely coexist in clouds and fogs. The species with the lowest concentration is the limiting reactant, and is rapidly depleted inside the cloud or fog layer (Pandis and Seinfeld, 1989b, cited in U.S.E.P.A, 1996a, pg. 3-27). In heavily polluted atmospheric water, metal catalysed oxidation of S(IV) by O2 may contribute significantly to the formation of sulphate in the liquid phase. In such situations, oxidation by O2 may be more important than oxidation by hydrogen peroxide.
2+ 3+

(2.12) (U.S.E.P.A., 1996a) This reaction is pH dependent, occurring at different rates depending on the pH regimen and the ionic strength of the solution. Nitrogen dioxide can react with HSO 3 - to oxidise S(IV) into S(VI).

51

(2.13) (U.S.E.P.A., 1996a) Pandis and Seinfeld (1989b, cited in U.S.E.P.A., 1996a, pg. 3-55) reported that in fogs occurring in polluted urban areas with high NO2 concentrations, this reaction could be a major pathway for the S(IV) oxidation if the atmosphere has enough neutralising capacity, i.e. high NH 3 (g) concentrations. The conversion of SO2 to sulphate in aerosols at 90% relative humidity can contribute perhaps 10% of the total sulphate production (90% due to gasphase oxidation of SO2 with OH). At higher relative humidities and/or lower temperatures the aqueousphase contribution would be expected to increase. As in the case of dissolved sulphur, aqueous-phase reactions of nitrogen dioxide may occur. A second pathway for the formation of nitric acid (the first being the gas-phase reaction) is the aqueous-phase reaction sequence:

(2.14) (2.15) (2.16) (U.S.E.P.A., 1996a) Once nitric acid has been formed, its reaction with ammonia in the gas phase may lead to the formation of particulate ammonium nitrate. Nitric acid may also react with particles containing sodium chloride or calcium carbonate; releasing HCL or CO2 and forming sodium nitrate or calcium nitrate, both of which may remain in the particle. Ammonia (NH3) will preferably react with H2SO4 to form (NH4)2SO4 and only if enough NH3 is available, the reaction with HNO3 will take place, forming NH4NO3. Thus, the role of ammonia is to neutralise the sulphates and nitrates in the atmospheric system. As SO2 emissions are reduced (consistent with some urban trends) and less H2SO4 is produced, more NH3 is available to react with HNO3, thus explaining the increasing proportion of the nitrate species in urban polluted aerosols. Chow et al (1994) published the average chemical composition of urban aerosols measured in the Burbank area, within the Los Angeles metropolitan area. They reported their data in several tables showing the chemical species concentration in g/m3. The table below displays an adaptation from their data into percentages.

52

Table 2.1 PM2.5 and PM10 Chemical Composition in Burbank, California 1987
Species Sulphate Nitrate Organic Elemental Carbon Ammonium Other Total Mass (g/m3) PM2.5 Percentage by Mass Summer Autumn 20.5% 4.8% 12.0% 28.1% 21.4% 25.0% 5.2% 8.1% 10.1% 8.3% 30.8% 25.7% 42.6 (g/m3) 78.3 (g/m3) PM10 Percentage by Mass Summer Autumn 14.7% 4.8% 14.7% 27.2% 17.7% 23.3% 4.3% 7.8% 6.9% 9.4% 41.7% 27.5% 72.3 (g/m3) 94.8 (g/m3) (Adapted from Chow et al, 1994)

2.3 Deleterious Effects


Depending on their chemical composition, which can vary from simple salts to strong acids, PM10 can have several effects on human health. They may promote a reduction in pulmonary function and at very high concentrations, they may cause pulmonary cancers and thus premature death of individuals. Specifically, suspended particles can have any one of the following effects: Cause toxic responses because of their inherent physical and chemical properties. Interfere with one or more of the functions in the respiratory system. Act as a vehicle to toxic substances that may be absorbed or adsorbed in the particle. In conjunction with other contaminants, such as sulphur dioxide, they may cause unsettled breathing, reduction of pulmonary volume and breathing difficulties along with irritation of the respiratory track. In addition, a positive correlation between concentrations above 150 g/m3 of PM10 (the Mexican Air Quality Standard) and mortality has been established. It has been shown that for every 10 g/m3 increment above the standard, the daily mortality rate due to the exposure of these particles increases by 3% (DDF, 1996). Apart from the health aspects, PM 10 have negative impacts on the environment by reducing visibility. This effect emanates from the scattering properties of the particles that, because of their size, interact within the visible light range of the spectrum. The overall negative effect of PM10 on visibility is augmented in the presence of high relative humidity, especially in the case where aerosols are composed of either the sulphate or nitrate species. Particles in the PM2.5 range are responsible for most of the light absorption and scattering, as normally these interfere with the visible light spectra and because they tend to be associated with the above compounds. In addition to the scattering of light, visibility is impaired by the absorptive properties of nitrogen dioxide (N02), a precursor of secondary particulates

53

(Yocom et al, 1986). Because of their inherent characteristics, in terms of human health and environmental hazards, PM10 's have been considered good indicators of atmospheric quality. They have now replaced the Total Suspended Particles (TSP) parameter as a criterion pollutant in the legal framework of Mexico and the United States, among other countries (Cicero-Fernndez et al, 1992).

2.4 Removal Mechanisms


The processes affecting the residence time of particles in the atmosphere can be quite complex and depend on both the physical and chemical characteristics of the aerosols. Generally, smaller particles will tend to coalesce and form new larger particles that can either settle faster or become more likely to activation as cloud condensation nuclei (CCN). This process, called coagulation, is especially important for the smaller particles, as it will shift the size distribution towards the larger sizes (Warneck, 1988). The process therefore involves nuclei particles growing into the accumulation mode, where they can be removed via wet deposition (as CCN) or shift into the coarse mode. Larger particles are more efficiently removed, as they are susceptible to both dry and wet deposition. Atmospheric particles are subjected to several forces that affect their movement in the troposphere. Depending on their size, their displacement is controlled either by Brownian motion or by the gravitational field of the earth. Brownian motion results from the constant bombardment of molecules from the surrounding fluid on the particles, while gravity depends on the mass and gravitational acceleration. The importance of each can be observed by comparing the distance travelled by a particle as a result of Brownian diffusion (diffusion caused by Brownian motion) and gravitational settling. Over one second, a 1 m particle diffuses a distance of about 4 m, while it falls about 200 m under gravity. On the other hand a 0.01 m particle will diffuse around 20 m and fall 4 m (Seinfeld and Pandis, 1998). It can be inferred from this information that Brownian diffusion is more significant in the smaller particles and that gravitational settling is important for the larger ones. Removal by coagulation mainly affects aerosols smaller than 0.1 m in diameter. In the removal as in the formation processes, particles coalesce to form new larger particles that are more efficiently removed from the atmosphere. Dry deposition is defined as the transport of gases and particles from the atmosphere to the surface in the absence of precipitation (Seinfeld and Pandis, 1998). It is mostly controlled by atmospheric turbulence, physico-chemical properties of the particles and by the nature of the surfaces. The size, shape and density of the aerosol will determine,in part, the deposition velocity. Particles are transported to the ground by turbulent diffusion and in the case of coarse aerosols, this process is enhanced by gravitational settling.

54

Turbulence intensity is largely dependent on the lower atmospheric stability and the surface roughness. During the daytime, it is typically large and it therefore increases the possibility of material being deposited onto the surface. In the night, however, vertical stratification near the surface reduces the intensity and the vertical extent of turbulence, hence reducing the scale of dry deposition. Before being deposited at the surface, particles pass through a layer of stagnant air just adjacent to the ground, called the quasi-laminar sublayer. Transport across this sublayer, by the falling particles, occurs either by Brownian diffusion or by sedimentation. Particles smaller than 0.05 m are transported through by Brownian diffusion, while the larger particles posses inertia and cross by impaction. A characteristic minimum in the deposition distributions is located in the accumulation mode particles. The main reason being, that there are no effective mechanisms to transport these particles across the quasi-laminar sublayer. Brownian motion ceases to be effective on particles greater than 0.05 m in diameter and inertial impaction is only important in particles larger than 2m. Deposition of aerosols larger than 10 m is dominated by gravitational settling, especially since the setting velocity increases with the square of the particles diameter (Seinfeld and Pandis, 1998). Taking into account all this information, it can be stated that dry deposition is an effective removal mechanism for nuclei (< 0.1m and in this case smaller than 0.05 m) and coarse mode (> 2 m) particles. The only deposition mechanism that actually removes accumulation mode particles is wet deposition, defined as the natural process by which material is scavenged by atmospheric hydrometeors and is consequently delivered to the surface (Seinfeld and Pandis, 1998). Removal via this pathway occurs in two regions, inside and below clouds, with the former being ten times more efficient (U.S.E.P.A., 1996a) Scavenging of aerosols in the clouds arises mostly by their activation into CCN. This occurs in the cloud formation process, with the particles serving as the material onto which water vapour condenses. Supersaturation ratios in the order of 400% are necessary for H2O vapour to condense on its own, while much lower and environmentally realistic ratios are required when a particle is present as nuclei (in the order of 1% or less, Seinfeld and Pandis, 1998). Although, with increasing relative humidity, all particles will contain a certain level of liquid water, only particles larger than 0.2 m in diameter will become activated and grow into cloud droplets (Warneck, 1988). The larger and more soluble particles will be activated first, while the some of the smaller will follow after the cloud (or air parcel) rises towards higher altitudes and the level of supersaturation increases. In continental clouds almost all particles larger than 0.4 m are expected to undergo transformation into a drop (nucleation scavenging). Smaller aerosols, around 0.2 m, are not efficiently scavenged and will remain in the cloud as interstitial aerosol. These size particles may only be removed via collisional capture with other cloud drops. Below cloud processes occur mainly by collisional impaction, with the falling drop incorporating particles in its trajectory to the surface.

55

The amount of particles removed from the troposphere is determined by the washout ratio (Wr ). Under the assumption of a vertically uniform concentration of aerosols, and that there are no other factors limiting collection (such as solubility), it relates the concentration of material in surface level precipitation to the concentration of the material in surface level air. With this relation, the vertically averaged scavenging ratio, , can be defined as a function of the wet deposition velocity, Vw.

Where and

(2.17) is the vertically averaged scavenging rate and h is the reference,

(2.18 where Wr is the washout ratio and P0 is the precipitation intensity height (U.S.E.P.A, 1996a). Together, in and below cloud scavenging (wet deposition) are the dominant processes removing particulate matter from the troposphere. One of the other factors affecting wet deposition is the solubility of the particles, whether they are hydrophilic or hydrophobic. Secondary particles composed of trace gases such as SO2 and NO2 have a different solubility and their incorporation into cloud water may be dissimilar to the idealistic representation presented before. Sulphur dioxide, for example, is moderately soluble in water, but decreases with increasing acidity. In-cloud reactions may transform the gas into different species that are more soluble, such as the S(IV) agglomerates. By contrast, nitrogen oxides have low solubility but their transformation into HNO3 allows them to dissolve in water and become subject to precipitation as part of rain (U.S.E.P.A., 1996a). Other pollutants have very low solubility in water and cannot be removed from the troposphere via wet precipitation, for example carbon monoxide. To conclude, and from the discussion presented thus far, it can be stated that removal mechanisms are dependent on the physical and chemical properties of particles, in particular their size and level of solubility. The different mechanisms are therefore mainly controlled by these two attributes of the aerosol population. Coagulation removes particles in the smaller range, <0.1 m, while dry deposition is an efficient mechanism for particles in the two extremes of the spectrum, smaller than 0.05 m (nuclei mode) and larger than 2 m (coarse mode). Aerosols in the middle range, from 0.1 to 2 m (accumulation mode) may only be removed by wet deposition after being activated into CCN and then precipitated as rain. Wet deposition is probably the most efficient mechanism as it removes particles in the whole size spectrum.

56

In order to conclude the discussion on atmospheric aerosols, one more topic needs to be addressed. Once aerosols are deposited to the surface, they can be resuspended and re-entrained into the surface level air. Resuspension of particles into the troposphere occurs mainly by mechanical and / or wind generated disturbance. The two are quite distinct and have different effects. Wind on the one hand has the potential to be continuous and affect large areas, while mechanical disturbance is more localised and happens in shorter and more intense intervals. The size distribution of resuspended particles is usually described as being bimodal, with the first mode in the 2 5 m range and the second in the 30 60 m (U.S.E.P.A., 1996a). The former can be expected to consist of secondary particles, mainly formed by GTP, while the latter are probably composed of primary material from the crust of the earth. One of the characteristics of resuspended material is that the smaller particles are usually attached to larger ones (for example soil particles). These agglomerates represent the largest part of the mass of tropospheric particulates. Unfortunately, very little research has been conducted in this area and not enough information is available to establish a pattern or to make reasonable conjectures on their fate and origin in the atmosphere.

2.5 Conclusion
Atmospheric aerosols are a mixture of primary and secondary origin particles. Among their most important characteristics are their physico-chemical properties and their size spectra. Particles are divided into different modes that determine, in part, their behaviour, environmental hazards and fate in the atmosphere. The size distributions mentioned in the preceding sections aid in their analysis by agglomerating them into different size spectra that determine their formation and removal mechanisms. The chemical properties of the particles determine their source and, to a lesser extent, their removal processes. Aerosols composed of condensible species, which tend to be in the PM2.5 range, are secondary particles arising from chemical reactions in either the gas or aqueous phase. The level of solubility of these substances influences their incorporation into cloud water and therefore their wet precipitation potential. In the previous sections, the size distributions of atmospheric particles were discussed and it was concluded that most particles are smaller than 0.1 m while most of their mass is agglomerated in the larger sizes. The formation pathways were also discussed and it was revealed that the smaller particles, PM2.5, are mostly of secondary origin formed through GTP or aqueous phase chemistry and that the larger, PM10, have mainly primary origins. 57

The deleterious effects of these aerosols were also discussed and it was determined that they may cause premature death to humans and substantial hazards to the environment, especially in terms of visibility. Furthermore, the removal processes, important to atmospheric particles of environmental interest, were mentioned. It was highlighted that the larger and smaller particles are removed in a relatively efficient manner, while the middle size particles, in the accumulation mode, are only removed through wet deposition. To conclude, a short note on resuspension mechanisms of aerosols was introduced and emphasis was given to the lack of research in this area.

58

Ozone Formation
Chapter Three

3 Introduction
Ozone is a photochemical oxidant that produces several deleterious effects on the environment. In humans, it induces shortness of breath and discomfort on breathing deeply. A decline in competitive athletic performance and eye irritation may also occur upon exposures to high concentrations. On vegetation, it has been demonstrated that yield reduction may occur at cumulative exposure above a threshold value of either 30 or 40 ppb, which are typical background concentrations (PORG, 1997). Ozone is produced through a series of oxidation and photochemical reactions; the main mechanisms are detailed in the following sections of this chapter. The first part provides a brief historical perspective, a description of the general photochemical processes and the radiation wavelengths of interest in the troposphere, together with their respective energy yields. Next, the main ozone formation processes are discussed. A section describing the background photochemical formation mechanisms of ozone and a further one outlining the processes important in the more polluted urban atmosphere are included. The basic hydrocarbon-NOx reaction cycles are exemplified via the oxidation mechanism of an alkane. Moreover, the sources of photochemical ozone and its precursors are provided throughout the chapter, and a section is included to explain the major tropospheric sinks of radical species, nitrogen oxides and ozone. One further section discusses the importance of the VOC/NOx ratio in the troposphere. The chapter concludes with an overview of the photochemical climates of cities.

3.1 Historical Perspective


Air pollution in urban areas has been historically typified by high atmospheric concentration of sulphur compounds (sulphur dioxide and sulphates) and particulates, resulting mainly from the combustion of coal and high sulphur containing fuels. Cities with this type of pollution are often situated in colder climatic regions, where power generation and domestic heating are major sources of emission. A second type of air pollution appeared with the widespread use of petrol as a motor fuel. Automobile exhaust pollution was first discovered in 1915, but it was not until 1945 that the first urban air pollution problem acknowledging the contribution of automobile emissions was recognised in Los Angeles, California, USA (Seinfeld and Pandis, 1998). This type of pollution occurs in almost every large urban agglomeration, including Mexico City and Greater Manchester. This second type of air pollution includes, among other contaminants, photochemical oxidants and its precursors and reacting atmospheric species. Photochemical smog results from the reactants and products of complex chemistry, that occurs when sunlight irradiates an atmosphere loaded with organic gases and oxides of nitrogen (NOx = NO+NO2). 61

Generally, photochemical smog is produced at high temperature and bright sunlit locations, but as the chemical bonds of certain compounds vary in strength, it can also occur in colder and relatively darker climates. The main primary pollutants involved in photochemical smog are nitric oxide (NO) and organic compounds, mainly volatile organic compounds (VOCs). These are readily converted to secondary pollutants that in turn have a further deleterious effect on the environment, ozone (O3) being one of the most important. The role of ozone in the atmosphere is twofold; in the stratosphere it is beneficial as it filters out harmful ultraviolet radiation from the sun, while in the troposphere it is a malignant oxidant when found at higher than background concentrations. Stratospheric ozone results from very complex chemistry involving the division of oxygen molecules by lightening and further reaction with other oxygen molecules. Although its depletion is a problem, it lies outside the scope of the present study. Tropospheric ozone on the other hand, results from the oxidation of VOCs in the presence of NOx and sunlight at certain wavelengths. The photochemical reactions involved in the formation and destruction of ozone are discussed in the following sections.

3.2 Basic Photochemistry


Photochemical reactions are initiated by the absorption of a photon in an atom, molecule, free radical or ion. The energy of one photon of frequency v is equal to hv, where h is Planks constant and the product hv is in turn equal to hc/ so that:

(3.1) (Seinfeld and Pandis, 1998) h = Planks constant (6.266 x10-34 Js), c = speed of light in a vacuum (2.99 x 108 ms-1), and = wavelength (in nm). To express the energy of a photon, , in moles of a substance, it is necessary to multiply hv by Avogadros number (6.022 x 1023 molecules per mole).

in KJ mol-1 (3.2)

(Sein The typical range of wavelengths and energies of interest in atmospheric chemistry lie in the visible and near the ultraviolet region of the spectrum. The corresponding ranges and their energy values are given in the Table 3.1.

62

Table 3.1 Typical Atmospheric Wavelengths and Associated Energies Name Red Orange Yellow Green Blue Violet Near Ultraviolet Vacuum Ultraviolet Typical Wavelength (nm) 700 620 580 530 470 420 400 200 200 50 Typical Range of Energies (kJ mol-1) 170 190 210 230 250 280 300 600 600 2400 (Seinfeld and Pandis, 1998) Photochemical reactions may occur when the energy of the incoming photon exceeds the binding energy of the particular chemical bond. The photon that is the reactant in the photochemical reaction is expressed as hv, consistent with equation 3.1. The primary step in any photochemical reaction is:

V i s i b l e

(3.3) Where A* is an electronically excited state of molecule A. The excited molecule A* may then undergo: Dissociation Direct reaction Fluorescence Collisional deactivation Ionisation (Seinfeld and Pandis, 1998) The former two reactions promote chemical change, while 3.6 and 3.7 return the molecule to its original state. (3.4) (3.5) (3.6) (3.7) (3.8)

63

3.3 Ozone Formation


In general, photochemistry begins with dissociation of the nitrogen dioxide molecule into NO and a single oxygen atom, and ozone is subsequently formed. The chemical mechanisms are detailed in the following section.

3.31 Basic Photochemical Reactions


Ozone formation occurs when an oxygen atom reacts with an oxygen molecule.

(3.9) (Seinfeld and Pandis, 1998) Where M is normally N2, O2, or another molecule absorbing the extra vibrational energy and thereby stabilising the O 3 molecule. All ozone is produced through this reaction. The source of atomic oxygen cannot be the O2 itself as only wavelengths exceeding 290 nm reach the troposphere and the oxygen molecules dissociate at wavelengths smaller than 200 nm. The main source of O3 forming single oxygen atoms is, therefore, the photolysis of nitrogen dioxide (NO2, which absorbs energy over the whole visible part of the spectrum, 400 nm).

(3.10) (Seinfeld and Pandis, 1998) Most NOx in urban atmospheres are however emitted as NO and it is only through its oxidation that NO2 can be produced, namely as a secondary pollutant. The basic conversion mechanism is the reaction of NO with oxygen via the following mechanism:

(3.11) (Seinfeld and Pandis, 1998) Once NO2 is formed, it photolyses to release an oxygen atom and a NO molecule. The O atom reacts with O2 to produce ozone. Nitric oxide, on the other hand, regenerates NO2 by reacting with O3.

(3.12) (Seinfeld and Pandis, 1998)

64

Reactions 3.9 to 3.12 represent a cycle of no net chemistry, establishing a photo-stationary state, normal to unpolluted background atmospheres. The high concentrations of ozone found in many places indicate that other formation pathways must exist. Ozone formed via reaction 3.9 may undergo photolysis to release either a ground state oxygen atom (O) or an excited singlet (O(1D)) oxygen atom, the type depending on the radiating wavelength.

(3.13)

(3.14) (Seinfeld and Pandis, 1998) The ground state O atom combines with O2 by reaction 3.9 to reform O3, leading again to a cycle with no net chemistry. The excited singlet, however, must react with another atmospheric specie as the direct transition from O (1D) O is forbidden (Seinfeld and Pandis, 1998). The most likely reactant species are N2 and O2, which remove the excess energy and quench the O(1D) to its ground state.

(k(298k)=2.6x10-11 cm3 molecules-1 s-1 ) (k(298k)=4.0x10-11 cm3 molecules-1 s-1 )

(3.15)

(3.16)

Occasionally, however, the singlet O (1D) atom collides with water vapour (H2O) and produces two hydroxyl radicals (OH). From this point onwards, the radical species will be indicated with a dot in front of the chemical formula (e.g. OH). (k(298k)=2.2x10 -10 cm 3 molecules -1 s -1 )

(Sein

(3.17)

At room temperature and about fifty percent relative humidity, as much as ten percent of the O(1D) produced reacts with H2O, mainly due to the concentration of H2O (104 ppm, 1%) and because the reaction rate constant (k) of reaction 3.17 is about a factor of ten larger than the quenching reactions with N2 and O2 (Seinfeld and Pandis, 1998). This percentage of hydroxyl radicals leads to an OH yield of approximately one molecule of OH per five O3 molecules photolysed.

65

In the background troposphere, this is the dominant source of hydroxyl radicals. Nonetheless, other sources exist and may become important under the appropriate conditions, some of which will be examined in the following section.

3.31 Sources of Radical Species


The photolysis of nitrous acid (HONO) can be a significant source of hydroxyl radicals during the early morning, especially in the urban boundary layer. HONO photolyses very efficiently at wavelengths less than or equal to 400 nm, to the extent that it can only be an important source of the OH after it accumulates during the night, in the absence of sunlight.

(3.18)

Nitrous acid is produced mainly through two reactions.

(3.19) (3.20)

Reaction 3.19 acts only as a reservoir of both NO and the OH radical while reaction 3.20 has a net formation of radical species as it can rapidly photodissociate back to nitric oxide and the hydroxyl radical (reaction 3.18). Another important source of OH is the reaction between the hydroperoxyl radical and nitric oxide.

(3.21) This source of hydroxyl radicals is important particularly throughout the day, since it provides a conversion pathway between the two radicals. Apart from its importance as a source of OH, it also oxidises a NO molecule into NO2 providing a conversion pathway between the two oxidation states, and therefore increasing the O3 formation capacity of the troposphere. The hydroxyl radical reacts with many of the atmospheric trace gases, mainly because it is unreactive towards oxygen and hence has the time and the concentration to promote other reactions. It reacts with carbon monoxide (CO) to catalyse its oxidation and produce carbon dioxide (CO 2 ) in the following cycle:

66

(3.22) (3.21) (3.23) (3.24) Net: CO+202+hv CO2+O3 (Seinfeld and Pandis, 1998) In this cycle, there is a net formation of O3 because the conversion of NO to NO2 is accomplished by the hydroperoxyl radical (HO2) rather than by ozone itself. This set of reactions can occur rapidly until one of the molecules is removed in a terminating reaction, such as in the case when OH and NO2 react to form nitric acid (HNO3); (3.25) or when the concentration of NOx is sufficiently low

(3.26) (Seinfeld and Pandis, 1998) Reaction 3.26 corresponds to an atmosphere loaded with VOCs and therefore has a high VOC to NOx ratio (explained further in section 3.35). The net formation of O3 in the troposphere is strongly dependent on the different oxidation pathways of nitric oxide, ultimately leading to the formation of nitrogen dioxide. The basic formation process is through reaction 3.11, although under most tropospheric conditions, this reaction is unimportant and the predominant mechanism is reaction 3.12. This process however, does not produce ozone and instead it consumes it. Several reaction cycles, like the oxidation of carbon monoxide and that of many organic species, readily oxidise the NO molecule with a subsequent net formation of O3. In most instances, trace gases in the troposphere react with hydrogencontaining free radicals (HOx = OH, HO2, H) to convert NO to NO2, with ozone formed subsequently via the photolysis of NO2 and reaction 3.9. As explained above, O3 photolyses (at 320 nm) to produce the singlet oxygen atom, which in turn reacts with water to produce the hydroxyl radical. This process is disturbed when the O3 concentration is reduced by reaction 3.12, typical of a NO rich, urban atmosphere. Under this condition, the photolysis of formaldehyde (HCHO) or its reaction with the hydroxyl radical dominates the production of HOx.

67

(3.27) or (3.28)

(3.29) (3.30) (Seinfeld and Pandis, 1998) The hydroperoxyl radical (HO2) then reacts with NO to produce NO2 through reaction 3.21 and subsequently to form ozone via reactions 3.10 followed by 3.9. Under the photolytic cycle, two NO NO2 conversions are accomplished and two OH are created, while in the second case only one conversion is promoted and the OH is regenerated. The hydroxyl radicals can then react with CO to form CO2 and O3 in the chemical cycle described above (reactions3.21 to 3.24).

3.32 Photochemistry of the Background Troposphere


The chemistry of methane (CH4) is of primary importance in the background troposphere, as it is the principal hydrocarbon species. It reacts with the OH to form the methylperoxy radical (CH3O2), which in turn may react with other pollutants, preferably with NO and HO2 but also with NO2, and other organic peroxy radicals (RO2). When it reacts with NO, the complete cycle has a net formation of HCHO, O3 and H2O: (3.31) (3.32) (3.33) (3.21) (3.34) (3.9) Net:

CH4+402+2hv

HCHO+203+H2O
68

(Seinfeld and Pandis, 1998

In total, two molecules of ozone are formed from each CH4 and the oxidation of HCHO leads to the formation of another O3 molecule (as illustratedabove through reactions 3.27 to 3.30 followed by 3.10 and 3.9). When the methyl peroxy radical reacts with HO2, it creates methyl hydroperoxide (CH3OOH), which can either photolyse or react with an OH.

(3.35) (3.36) (Seinfeld and Pandis, 1998) The methoxy radical (CH3O) may then react with O2 via reaction 3.33 and follow the cycle above. On the other hand, if it reacts with the hydroxyl radical; (3.37) or (3.38) and (3.39) (Seinfeld and Pandis, 1998) The lifetime of methyl hydro-peroxide in the troposphere is calculated to be in the order of two days, in reality, therefore, it may provide a sink for radicals through its wet and dry deposition. Logan et al (1981) (quoted in Seinfeld and Pandis, 1998) calculated that the reaction of the methyl peroxy radical with NO dominates over that of the hydroperoxyl radical, when the NO mixing ratio is more than thirty parts per trillion, while for lower ratios the latter is preferred. The oxidation of methane is important not only because it generates O3, but also because some of the intermediate species in the cycle above may act as reservoirs of NO2. The methyl peroxy radical may react with NO2 to create methyl peroxy-nitrate (CH3OONO2), a stable molecule that may live up to 2 days in the upper troposphere.

(3.40) (Seinfeld and Pandis, 1998)

3.33 Photochemistry of the Urban Troposphere


In all but the most pristine areas, and especially in urban environments, the chemical reactions of organic species, coupled with the anthropogenic

69

NOx emissions, drive the formation mechanisms of ozone. Most of the reactions referred to in the previous sections are important in the rural and background troposphere, where VOC and NOx emissions are comparatively low. The chemistry of the urban environment is significantly more complicated due to the presence of VOCs of various classes (alkanes, alkenes and aromatic hydrocarbons). In the background troposphere, the chemistry of methane and its degradation products (like formaldehyde) represent most of the chemical mechanisms available for the production of ozone and hence, for the NO NO2 conversion pathways. In the urban environment, however, the oxidation of non-methane organic compounds (NMOC) in the presence of NOx dominates the production and conversion mechanisms. Their oxidation, in most cases, is driven by the reaction with the hydroxyl radical to produce organic radicals (R), organicperoxy radicals (RO2) and further degradation products (PORG, 1997). As it can now be inferred, the key reactive specie in urban ozone formation is the hydroxyl radical. It is regenerated through the major chemical cycles (sometimes in multiples) and provides several conversion pathways for the NO NO2 (through the intermediate production of the hydroperoxyl radical) by reacting with various classes of VOCs.

3.331 Chemistry of NOx and Organic Compounds


It can be generally stated that photochemistry starts with the dissociation of nitrogen dioxide via reaction 3.10 and the subsequent ozone formation through reaction 3.9. The efficiency of this system is restricted by the different pathways that exist to oxidise nitric oxide to its photoactive state, NO2. As stated earlier, the OH plays an essential role in the NOx conversion process. Apart from the mechanisms already mentioned (important in the background troposphere) it reacts with hydrocarbon species (RH) to produce alkoxy-radicals (and further oxidation states) that eventually lead to the formation of the hydroperoxyl radical and its reaction with NO to form NO 2 . (3.41) (3.42) The organic peroxy radicals (RO2) react with NO to form NO2 and other free radicals.

(3.43) or (3.44)

70

They may also react with NO 2 to form organic peroxy nitrates via:

(3.45)

Reaction 3.43 is very similar to the HO2NO mechanism, but instead of regenerating the OH, it creates the alkoxy radical (RO). The most common fate of RO is reaction with O2, leading to HO2 formation and a carbonyl compound.

(3.46)

A good example of this mechanism is the reaction of the simplest alkoxy radical, methoxy (CH3O), as it reacts with O2 in reaction 3.33 to form HCHO and HO2. Reaction 3.45 occurs at an order of magnitude slower than the photodissociation reaction of NO2 (reaction 3.10), but depending on the organic group, it can form very stable molecules that can persist in the troposphere for relatively long periods. The reaction of aldehydes with the OH, especially of acetaldehyde (CH3CHO), is important as it leads to the formation of peroxy acetyl nitrate (PAN = CH 3 C(O)O 2 NO 2 ), an important reservoir for both HOx and NOx.

(3.47) (3.48) (3.49) (3.50) (3.51) (Seinfeld and Pandis, 1998) PAN is an important NOx reservoir, as it is not very soluble in water and it is relatively inert at temperatures typical of the upper troposphere. Thermal decomposition by the reverse of reaction 3.49 is its main loss process and at the same time, it replenishes the NO2 molecule and produces a peroxyacyl radical. The concentration of PAN depends on the temperature of the location (it becomes more stable at lower temperatures) and on the local source of NO2 and NO, the latter because of the competing mechanism (reaction 3.50) that removes

71

the peroxyacyl radicals from the PAN system.Its lifetime is strongly dependent on temperature and ranges from around thirty minutes at 298 K to about eight hours at 273 K. At the temperatures of the upper troposphere, PAN may live up to several months, a situation that may encourage other degradation processes (such as photolysis or reaction with OH). During the night, the thermal decomposition of PAN in an atmosphere loaded with NO (e.g. 10 ppb) can initiate reaction 3.50 to convert NO into NO2 and produce a methylperoxy radical (CH3O2) (Seinfeld and Pandis, 1998). This in turn may convert another NO to NO2 and produce a formaldehyde molecule and a hydroperoxyl radical (reactions 3.32 and 3.33). Finally, the hydroperoxyl radical may react with NO to regenerate the OH and convert one more NO into NO2 via reaction 3.21. These type of mechanisms are termed propagation reactions as they regenerate the OH loading in the troposphere and maintain the photochemical cycle necessary to reach the high concentrations found in many urban agglomerations. The general mechanism in urban polluted environments is the reaction of a hydrocarbon molecule with the hydroxyl radical to produce a carbonyl molecule (RCHO) and two NO NO2 conversions. The hydroxyl radical is then regenerated at the end of the specific set of reactions. The generic chemical mechanism outlined above (using R to denote organic compounds) describes the oxidation of an alkane. The hydroxyl radical attack on other VOCs and their subsequent oxidation generally occurs by similar mechanisms. The rate of oxidation of VOCs is governed by the tropospheric concentration of the HOx radical species, which in turn is controlled by a balance between their sources and sinks.

3.34 Tropospheric Sinks of Ozone and Its Precursors


As has been noted throughout the chapter, the concentration of ozone is dictated by the amount of HOx available to oxidise VOCs in the troposphere. The principal removal mechanisms of HOx are through reaction with itself (mutual termination);

(3.26) (3.52) (3.53) or through their reaction with the NOx species.

72

(3.25)

(Weinstock, et al, 1980 Reaction 3.25 is the predominant removal mechanism for HOx and NOx, especially in urban polluted atmospheres, where the concentrations of both species are relatively high. The former set however, becomes significant only when the concentration of HOx is very high and / or the concentration of NOx is relatively low or close to zero. The removal efficiency of reaction 3.44, relative to the other pathway (reaction 3.52), is strongly dependent on the size and structure of the RO2. Taking into account the peroxy radical population, reaction 3.44 explains only one percent of the overall reaction mechanism. It is therefore believed to be of minimal importance under normal tropospheric conditions. The major HOx sinks (reactions 3.25 and 3.26) act in reality as reservoirs of HOx species (and of NOx species in the case of reaction 3.25) mainly because the possibility of regeneration through chemical decomposition or photolysis often exists. Nitric acid (HNO3) may photolyse to regenerate NO2 and OH through; (3.54) and hydrogen peroxide may also photolyse via:

(3.55) (Weinstock, et al, 1980 It is reservoirs like these that allow HOx to be available for reaction with organic and other species in the troposphere and therefore to generate photooxidants, like ozone. As in the case with HOx, the principal loss of NOx is in the form of nitric acid (HNO3). This molecule can be formed through several mechanisms. The principal reactions involved in the formation of HNO3 through NOx are that of OH with NO2, 3.25, through the night time species (N2O5) and water, and the reaction of the nitrate radical and trace organic compounds. During the night, NO2 is slowly converted to the nitrate radical (NO3) by reaction with ozone; (3.56) and further reaction with NO2 leads to the formation of N2O5, a reaction that rapidly reaches equilibrium between the two species. (3.57)

73

N 2 O 5 then reacts with water to form two molecules of nitric acid.

(3.58) (PORG, 1997) Similarly, the nitrate radical may also react with several organic compounds, including alkanes, alkenes, aromatics and sulphur containing organics.

(3.59) During the day, however, the nitrate radical is photolysed very efficiently leading to the regeneration of NO2 and an O atom, which rapidly generates O3. (3.60) Consequently, its importance is limited to the night time chemistry. It provides a reservoir for NOx and the reaction between NO3 and NO provides another conversion pathway (NO NO2).

(3.61) (PORG, 1997) In any case, it provides a further removal mechanism of NOx and for organic molecules, which in general tend to have long atmospheric life times. As in the case of HOx, nitrogen oxides have atmospheric reservoir species that allow them to accumulate and increase their potential to create ozone and other photochemical pollutants. Included in these are PAN, nitrate radical, N2O5, and HNO3 itself. The principal removal mechanisms of ozone are the titration reaction with NO (reaction 3.12) and photolysis via reaction 3.14. The former reaction removes both pollutants and promotes one NO NO2 conversion pathway, while the latter is only a sink when the O(1D) atom reacts with H2O to form two hydroxyl radicals via reaction 3.17 otherwise the O(1D) is just quenched back to O immediately reforming O3. The only other major sink available in the gas phase is through reaction with HO2 via: (3.62) (Seinfeld and Pandis, 1998) Another important ozone sink is through the aqueous phase. Although ozone is relatively insoluble in water droplets, the presence of HOx radicals

74

in solution can lead to a further removal mechanism. In solution, HO2 ionises to form an O2- ion;

(3.63) which can react with ozone and remove it via:

(3.64) (PORG, 1997) The net effect on O3 loss through this manner is further enhanced since the scavenging of HO2 in clouds is usually in competition with the gas-phase reaction of HO2 and NO. This process is thought to account for about 20% of the total free tropospheric ozone removed (PORG, 1997).

3.35 VOC to NOx Ratio (VOC/NOx)


The relative abundance of NOx and VOCs in the troposphere can play an important role in the formation of ozone. Both species are in competition for the radical molecules and as such the predominance of one over the other determines how fast or slow, and to what extent, ozone is produced. In general, radical reactions with NOx are much faster (a factor of five) than those with hydrocarbons (Seinfeld and Pandis, 1998). A high VOC/NOx ratio indicates the predominance of VOC, a condition termed NOx limited. Where NOx are more abundant, however, the condition is denominated VOC limited. The whole system is determined by chain terminating or propagating reactions, the former being one in which more radical species are removed than created, while the latter where more are created than destroyed. The effect of a change in concentration of either species is represented by the chain length, the larger the chain the larger the potential for ozone formation. A reaction that promotes the formation of ozone, that is, produces more radicals, is considered to increase the chain length, while one which destroys or removes radical species from the system is said to reduce the potential. Under a VOC limited condition (low VOC/NOx ratio), the main chain propagating reaction is 3.41, while the principal terminating reaction is 3.25. In this case, a decrease in VOCs (RH) would decrease the chain length and the overall result would be a decrease in the rate of ozone formation. By contrast, a decrease in NOx (NO2) would provoke an increase in ozone production. This occurs because the terminating reaction depends on the availability of NOx; the less available, the larger the chain length and the higher the potential to produce O3 (PORG, 1997).

75

In contrast, when the concentration of NOx is the limiting factor (high VOC/NOx ratio), the dominant chain terminating reactions are 3.21 and 3.43, while the main propagating mechanisms are 3.26 and 3.52. Under this condition, a decrease in either species would reduce the chain length leading to a decline in ozone production. Ozone reduction in this case is best achieved by reducing the concentration of both VOCs and NOx. In a high VOC/NOx ratio condition, NOx are constantly being depleted from the system (as they react faster with HOx) up to the point that it reaches a concentration where the reaction between peroxy species becomes important (reaction 3.26) (PORG, 1997). In general, a reduction in the production of ozone is best achieved by a decrease in the VOC concentration under VOC limiting conditions and by a reduction in the concentration of NOx under a NOx limiting situation. This can be observed in an ozone isopleth, Figure 3.1: Figure 3.1 Ozone Isopleth (adapted from Seinfeld, 1986)

Ozone Isopleth, Organic/NOx Chemistry


0.60 0.50

NOxppm as NO2

0.40 0.30 50 ppm 0.20 0.10 12 ppm 0.00 0.25 0.50 0.75 1.00 1.25 1.50 1.75 2.00 2.25 2.50 30 ppm 40 ppm

VOC ppmC

3.4 Conclusion
Ozone formation is dictated by the availability of its precursors in the atmosphere; the relative abundance of HOx and organic species is especially important in urban centres where concentrations can be very high. The ozone climate of a city is determined, in part, by its geographical situation, the pool of precursors, and the type and quantity of pollutant emissions. The geographical location affects the formation of ozone in many different ways. Certain locations are more prone to photochemical formation

76

and transformation than others. In the case of ozone, formation is favoured at high altitude and low latitude locations, where irradiation levels are higher and the amount of water in the air (relative humidity) is greater. As stated, the level of insolation is important, as ozone is a photochemical pollutant and requires solar irradiation at certain wavelengths; the larger the amount of sunshine at relatively short wavelengths, the greater the potential formation of ozone and other photochemical oxidants. This again is favoured at low latitudes where the average amount of irradiation at the surface will tend to be higher throughout the year. Additionally, locations prone to high-pressure systems will be inclined to accumulate pollutants (at times of thermal inversions and light winds) and will therefore allow this pool of pollutants to interact more thoroughly and sometimes at higher concentrations. Thermal inversions, typical of high-pressure systems, trap pollutants close to the surface, with the inversion layer acting as a lid over the boundary layer. The type of pollutants emitted into the troposphere is crucial in determining the ozone formation potential. As noted earlier, photochemistry begins with the photodissociation of nitrogen dioxide, the oxidised state of NOx. Most NOx in the troposphere are emitted as NO, while NO2 is a secondary pollutant formed by the oxidation of nitric oxide. The total amount of conversion mechanisms of NO NO2 determines the ultimate amount of NO 2 produced and thus, under most tropospheric circumstances, the total amount of ozone produced and consumed. The conversion pathways are, in turn, controlled by chemical cycles that, in most cases, are initiated by reaction with the hydroxyl and hydroperoxyl radical (HOx). The whole system, therefore, requires NOx (which will most likely be in the form of NO, thus the pathways for its conversion to NO2 are essential), HOx, and sunshine. The large concentrations of O3 found in most urban centres require, in addition to these, the presence of organic compounds to catalyse their formation through oxidation and resulting conversion mechanisms of NO NO2. Thus, the presence of VOCs is also essential in the formation of ozone, especially at locations with elevated concentrations. In essence, the formation of ozone occurs through the oxidation of VOCs in the presence of NOx and sunlight. As with any other pollutant, ozone concentrations are largely dictated by its sources and sinks. The sources are, in the most part, petrol fuels, motor exhausts and industries emitting NOx and hydrocarbons. A large part of the constitution of petrol fuels is dominated by hydrocarbons, especially alkenes and aromatic hydrocarbons. Both of these species are highly volatile and promote ozone formation through oxidation, which in turn creates and regenerates HOx. In addition, motor exhausts are loaded VOCs and produce large quantities of NO. In the case of industrial emissions, the generation of power through natural gas fuels has replaced the highsulphur-content emissions with NOx

77

and hydrocarbons. The large amounts of NOx and VOC emissions found in most urban agglomerations, coupled with the oxidative nature of the atmosphere, creates a medium prone to photochemical smog. At locations with elevated O3 concentrations, the sources are mainly anthropogenic while the sinks are mostly naturally occurring reactions in the atmosphere (both chemical and physical). Cities with high emissions of NOx and VOC and that are located in places promoting photochemical oxidation (through the characteristics described above) will have climates favouring ozone formation.

78

Metropolitan Areas
Chapter Four

4 Introduction
Urban agglomerations have inherent environmental problems that may result in poor quality of life to their inhabitants and to those residing in areas surrounding them. Important among the factors that may contribute to environmental exacerbation are the geographic location of the area and therefore its climatology and meteorology. The organisation of the society that dwells in the area is also important, as this will shape the activities that occur within the agglomerations and the behaviour of their population. The regions in which the Mexico City Metropolitan Area (MCMA) and the Greater Manchester Area (GMA) are located differ greatly, but they experience similar problems. The air quality of both areas is relatively poor, especially in relation to the surrounding countryside. The environmental problems that each face has lead their respective governments to implement emission control regulations and strategic programs aimed at reducing the concentration of the most common air pollutants. In the following sections, the areas of both urban agglomerations are described separately. The geographic factors affecting the concentration of atmospheric pollutants will be presented, including among others the location, climate and the meteorology of the regions. The different environmental regulations that each government has instigated will also be portrayed. In both areas, a monitoring network measuring the most common air pollutants has been installed and these will be described in terms of their geographic coverage and operation.

4.1 Mexico City Metropolitan Area (MCMA)


The area known to most Mexicans as Mexico City has been defined, for ecological purposes, as the area with a concentration of carbon monoxide (CO) greater than is naturally found in provincial areas. It now covers all of the Federal District (D.F.), parts of the State of Mexico and of the state of Tlaxcala, three different political entities (Williams et al, 1995). The megalopolis is one of the most cosmopolitan areas of the world, enveloping a very diverse and rich cultural background and a mix of social, political and environmental circumstances that make it a very interesting place to live in. Unfortunately, it is also heavily polluted, over populated and has become the subject of many research investigations that try to explain and promote the control of the many factors that contribute to both its positive and negative aspects. Environmentally, the area is affected by its geography and a combination of socio-political, demographic, and historical factors that not only affect the citizens but also the natural surroundings of the capital city.

81

4.11 Geographic Description


The MCMA lies at an average elevation of 2,400 meters (approximately 7500 feet) above the main sea level (msl) in a U shaped basin that opens to the North. Situated in the Anahuac Cordillera, it is surrounded by the mountains of the Sierras of Ajusco, Chichinautzin, Nevada, Las Cruces, Guadalupe and Santa Catarina, all of which form a physical barrier with a height of approximately 3700 m (12,000 ft) above msl. Two isolated peaks, the volcanoes Popocatepetl and Iztaccihuatl, reach elevations in excess of 5300 m (17,400 ft) on the eastern end of the basin. The MCMA is located in the central region of Mexico around the middle latitudes (1925N, 9910 W) and as such it is subjected to anti-cyclonic systems generated in the Golf of Mexico and the Pacific Ocean. These systems produce the ideal conditions for a very stable atmosphere, generating thermal inversions and inducing severe pollution episodes (DDF, 1996). The climatic conditions are mostly dominated by a semi-arid weather pattern with a mean annual precipitation dropping from about 900 to 400 mm as one travels from west to east of the city. The wet and dry seasons are fairly well defined, June to September and October to March respectively. The seasonal variation in temperature is small, while the diurnal range is large, with a mean around 16 degrees Celsius (Oke et al, 1992). The compounded effect of its geographic features contributes to make the MCMA a highly polluted metropolis and hence an unhealthy place to live. Some of the factors affecting the concentration of the air contaminants are summarised below: 23% less oxygen than the sea level average (because of its high altitude above the msl); Stagnation of air (due to the U shaped valley location); Very stable atmosphere (because of the anti-cyclonic effects); Low thermal inversions, especially during the winter when the ground is cold; Abundant solar radiation (which increases the photochemical effect of certain pollutants).

4.12 Factors Affecting Pollution Levels


In addition to the geographical circumstances that dominate the valley of Mexico, there are other factors affecting the high levels of pollution found in the metropolis. Among some of the most important and related patterns are socio-economic, demographic and historic variables, all of which must be taken into account when considering the environmental condition of the city. Socio-economically and demographically, the city attracts large amounts of immigrants (estimated to be around 10,000 each week) and hence has a very large growth rate. It is the centre of most of the economic activity in the country, as it houses a large proportion of the industrial, service and government 82

sectors of the nation, generating 34% of Mexicos GDP and 42% of its industrial revenue (IMP, 1994). It is one of the largest metropolis in the world, providing a home for more than 15.4 million inhabitants (INE, 1997). Historically, the capitals of most Latin American countries colonised by Spain during the 16th century developed in a pattern whereby such cities were the centre of all activities and all other urban and suburban agglomerations either depended on these or provided goods and services for them. The combination of all these, and other, factors has made the valley of Mexico an overcrowded city in almost every sense: in the number of inhabitants; the number of residential, industrial, commercial and service centres, and the amount of automobiles (more than 2.7 million cars of which it is estimated that 2.1 million operate every day, INE, 1997). The combined effect of these factors has produced a place where, in conjunction with its location and geography, the problem of air pollution has reached an unprecedented health and environmental hazard.

4.13 Environmental Controls


Environmental regulation in Mexico, like in many other parts of the world, stemmed from both the need to protect the inhabitants and their surrounding environment and from the public pressure to increase quality of life. The authorities governing the MCMA have experienced both of these pressures, the result of which has been the establishment of several government regulations, increased civil participation and intensive institutional research into the cause, effect and control of air pollution. During the present decade, the government of the Federation and of the city have implemented several control strategies to reduce the pollution levels of the MCMA. Several different programs have been instigated; some that require the approval of the civil society and others that have been placed without the need for a public referendum. Perhaps the most outstanding is the One Day Without A Car Program that was implemented in 1989. Under this program, the citizens of the MCMA are not allowed to use their automobiles for one day out of the working week, hence reducing both the amount of circulating traffic (by at least 20%) and the pollution associated with the automobile pool of the metropolitan area (DDF, 1990). In order for the government to implement such a program, the citizens had to approve it and a referendum took place to make its practice constitutional and therefore legal. In addition to this, the government has invested large amounts of resources into improving the overall environmental condition of the city. It implemented a contingency plan for episodes of severe pollution in which the industrial base of the capital city would have to reduce its production, or in some extreme cases halt it completely to reduce the emission of several contaminants. Along with the industrial requirement, the plan also stipulates one extra day without a car for the citizens.

83

This contingency has the purpose of reducing the concentration of the pool of pollutants to the established norm in the fastest way possible. Moreover, it forced the Mexican Petroleum Company (PEMEX) to reformulate leaded gasoline to reduce the lead content by more than 92%, and to increase the availability of unleaded fuels for new automobiles (post 1991). The automobile industry was compelled to install catalytic converters in all the new cars, in order to reduce the production of NOx and other exhaust gases, and to change the required petrol to unleaded. Included in these programmes was the complete phase-out of leaded petrol, a target that was achieved in 1997. The government also required PEMEX to reduce the content of sulphur in diesel fuel (reduced by 95%) as it is used extensively in public transportation and other sectors of the economy. The authorities closed one of the biggest coal-fired oil refineries (Refinera 18 de Marzo), situated to the north-east of the MCMA, as it was producing large amounts of contaminants in the area where prevailing winds disperse the pollutants throughout the metropolis. The government of the city and of the State of Mexico also established the norm of vehicular inspection every six months on all automobiles, hence requiring a high standard of engine function and reducing the pollutants emitted by badly tuned vehicles (DDF, 1990). As is quite evident from the programmes outlined above, most of the regulatory efforts are being focused around the automobile pool of the city, the reason being that it is stipulated that 75% of the pollutants come from the transport sector of the economy as shown in Figure 4.1 (DDF, 1996). Figure 4.1 Pollution Inventory by Sector, MCMA 1994 (DDF, 1996) Vegetation and Ground 12% Industry 3% Services 10%

Transport 75% The Secretariat of Public Health, through the Public Health Institute (INSP) is responsible for setting the legal standards of pollutants.

84

The norms that apply at present were set using the results of epidemiological, toxicological and exposure studies conducted in other countries. At the time when the standards were established, the country lacked the research facilities and human resources to undertake such analyses. Today, however, the facilities and resources are available and the secretariat is conducting epidemiological studies to validate the dose-response relationship between the different contaminants and the inhabitants of the city and to re-evaluate the standards in order to tailor them to meet the needs and requirements of the population within the MCMA (DDF., 1996). The norms that apply now are specified in two exposure periods, acute and chronic, and are presented in Table 4.1. Table 4.1 Mexican Standards
Pollutant Acute Exposure Chronic Exposure Average Concentration Max. Acceptable For the protection of and Time Frequency the susceptible population 0.11 ppm (1 hr.) 0.13 ppm(24 hrs.) 0.21 ppm (1 hr.) 11 ppm (8 hrs.) 150 g/m3 (24 hrs.) --260 g/m3 (24 hrs.) Once every 3 years --Once a year 0.03 ppm (annual mean) Once a year --Once a year --Once a year 50 g/m3 (annual mean) --1.5 g/m3 (3 month mean) Once a year 75 g/m3 (annual mean)

O3 SO2 NO2 CO PM10 Pb TSP

(Adapted from DDF, 1994) The National Ecology Institute (INE), in conjunction with the Department of the Federal District (DDF), operate the Automatic Atmospheric Monitoring Network (RAMA) and hence are in charge of studying and divulging the environmental information produced by the network. The INE is responsible for all the environmental audits in Mexico along with all the environmental matters of the federal government. The DDF established the General Directorate for the Control and Prevention of Environmental Pollution in the Valley of Mexico to co-ordinate the environmental efforts made by the Government of Mexico City and to administer the resources allocated to this entity. These two institutions carry out many other tasks but for the purpose at hand, these are the most relevant.

4.14 Automated Monitoring Network of the MCMA (RAMA)


The first automatic monitoring network in Mexico City was established in 1974 with a system consisting of 20 stations measuring several pollutants and meteorological parameters.

85

This system worked until 1980 when it was decided that it needed to be replaced as the information generated was not reliable and the equipment out of date. In 1984, it was substituted for the current system which later it became known as RAMA (Red Automtica de Monitoreo Atmosfrico or the Automated Atmospheric Monitoring Network). This system became operational in October of 1986, and was finally completed in 1991. Since then, most efforts have concentrated on the optimising and validating processes of the system and in widening the geographical coverage of the network. As with most automatic systems, RAMA measures the atmospheric concentration of the basic pollutants i.e. ozone (O3), carbon monoxide (CO), sulphur dioxide (SO2), nitrogen oxides (NOx), nitrogen dioxide (NO2), total suspended particles (TSP), and particulate matter under 10 micro-metres (PM10). In addition, all the monitoring stations are equipped to measure common meteorological parameters, i.e. temperature, relative humidity, and wind speed and direction. The whole system was designed and operates in accordance with the criteria set by World Health Organisations (WHO) in terms of the required number and location of the stations, and the parameters measured (DDF, 1993). The data from each of the different monitoring stations can be accessed via the internet at the following address: http://www.ine.gob.mx/calaire/espanol/mvm.html

4.141 Development of the Monitoring Network


The whole of RAMA was projected to be operational in three main stages; integration, extension and reinforcement, and consolidation. The first stage started in 1984 with 20 stations measuring the concentration of the most common pollutants (mentioned above). The second stage began in 1991 with the integration of seven new stations distributed throughout the urban area and including, for the first time, parts of the State of Mexico. PM10 analysers were introduced during this stage in five of the monitoring stations, providing coverage in the four cardinal points and the city-centre. The last part of the project commenced in 1993 with the installation of relatively sensitive equipment to monitor VOCs and Reactive Organic Gases (ROG). During this last stage, a quality assurance program was introduced to ensure the accuracy and comparability of the data. Over time, RAMA has become quite reliable and the data acquired through it is currently used to define environmental policies and strategies to control and prevent the pollution problems of the MCMA. The network it is one of the most extensive and comprehensive systems in the world (DDF, 1993). Table 4.2 depicts the size of some the monitoring systems operational in cities with a similar or greater territorial extension.

86

Table 4.2 Size Comparison of Monitoring Networks over the World City Mexico City Metropolitan Area Tokyo, Japan Los Angeles, Ca., USA Sao Paulo, Brazil Amsterdam, Netherlands Houston, Tx., USA No. 32 30 25 25 10 8 (DDF, 1993)

4.142 Atmospheric Monitoring System


The overall network is composed of three subsystems; the automated,meteorological and manual networks, each of which complement each other to provide a greater coverage. The first two monitor continuously and transmit the data to a central computer. The latter is checked once a week by personnel from the network

4.1421 Automatic Network


This section of the network is composed by 32 stations connected telemetrically to a central computer, where the data are analysed, checked and disseminated. The automated network monitors the most common atmospheric pollutants continuously throughout the year, 24 hours a day, 365 days of the year. It is integrated by equipment using the following modes of analysis: Table 4.3 Quantity of Monitors and Mode of Analysis (RAMA) Pollutant O3 CO SO2 NOx PM10 Total Hydrocarbon Quantity 19 21 21 8 5 5 Mode of Analysis Ultraviolet light photometry Non-Dispersive Infrared spectroscopy Pulsating fluorescence Chemiluminescence (NO and NO2 by subtraction) Attenuation via beta radiation Flame ionisation (DDF, 1994) 87

4.1422 Manual Network


The main purpose of the manual subsystem is to complement the automated network in the analysis of suspended particles and their elemental composition. It is made up of 19 stations, five of which have gravimetric analysers of PM10 while ten of them monitor heavy metals such as lead, copper, zinc, nickel, and cadmium via atomic absorption spectrophotometry. This network operates with a frequency of once every six days (24 hr. periods every 6 days, except during winter when the frequency may be intensified) in accordance with the specific criteria dictated by the WHO.

4.1423 Meteorological Network


This part of the network is composed of ten stations measuring the basic meteorological parameters, i.e. temperature, relative humidity, wind speed and direction. Additionally, it is complemented with a high tower equipped to measure high altitude wind and the conditions of the mixing layer. The instruments used to measure these parameters are described in Table 4.4.

Table 4.4 Meteorological Parameters and Mode of Analysis Parameter Wind Velocity Wind Direction Temperature Relative Humidity Temperature of the mixing height Mode of Operation Conventional weathercock Conventional anemometer Slim film capacitor Thermistor Polymer slim film capacitor Acoustic radar

(DDF,1993) This system is also responsible for providing the monitored information to the National Meteorological Service.

4.1424 Supporting Units


The network is rigorously monitored by a quality assurance program which calibrates, maintains, and certifies the stations (DDF, 1993). All the data is either ratified or rejected according to the specific requirements established by national and international organisations, such as the SSP, INE, and the WHO.

88

Through this program, it is possible to determine the overall efficiency of the network. In 1995, the system operated in optimal conditions transmitting accurate data 91.1% of the time. In the remaining 8.9% of the cases, the problems were related to modems, lost telephone lines, maintenance of the stations, and loss of energy, as shown in Figure 4.2 (DDF, 1995). Figure 4.2 Percentage Distribution of Faults (RAMA) Problems with the Automated Network
Modem Problems 6.0% Others 3.0%

Maintenance 25.4%

Loss of Telephone Lines 44.8%

Loss of Energy 20.9%

(DDF, 1995)

4.145 Metropolitan Index


The levels of pollution within the MCMA are communicated to the public in a homogenous unit called the IMECA (Indice Metropolitano de la Calidad del Aire or Air Quality Metropolitan Index). The index is formulated to have two main breaking points. The first indicating the Mexican Air Quality Standards (at a value of 100 IMECAs) and the other at the value where there is significant evidence that the concentration may pose a threat to human health (500 IMECAs). Two types of indexes are available, one for each of the pollutants and one to represent the general air quality of the atmosphere. The latter takes the value of the highest concentration (in IMECAs) of any of the pollutants on all of the stations, representing in this form the worst case scenario. The values are broken down into four different levels of pollution and are reproduced in Table 4.5.

89

Table 4.5 Representative Values of the IMECA IMECA 0 - 100 101 - 200 201 - 300 Air Satisfactory Not satisfactory Bad QualityEffects Favourable for all type of activities. Increased rate of small grievances. Increased intolerance to exercise and small grievances in persons with respiratory problems. Unhealthy symptoms, and intolerance to exercise by the general public (in good health). (DDF,1993) The values for the IMECA are formulated according to the algorithms presented in Table 4.6. Table 4.6 Algorithms Used to Calculate the IMECAs Pollutant Range in original units PM10 0-50 g/m 51-350 g/m3 351-420 g/m3 421-500 g/m3 501-600 g/m3 0-0.13 ppm 0.13-1.00 ppm 0-11 ppm 11-50 ppm 0-0.21 ppm 0.21-2.00 ppm 1-0.11 ppm 0.11-0.6 ppm
3

301 - 500

Very bad

Algorithm IMECA = (Conc. of PM10) IMECA = (1/2) x (Conc. of PM10) + 25 IMECA = (10/7) x (Conc. of PM10) - 300 IMECA = (10/8) x (Conc. of PM10) - 225 IMECA = (Conc. of PM10) - 100 IMECA = (Conc. of SO2) x 769.230769 IMECA = ((Conc. of SO2) x 459.770114) + 40.22989 IMECA = (Conc. of CO) x 9.090909 IMECA = ((Conc. of CO) X 10.256410) - 12.820512 IMECA = (Conc. of NO2) x 476.190476 IMECA = ((Conc. of NO2) x 223.463687) + 53.07264 IMECA = (Conc. of O3) x 909.090909 IMECA = ((Conc. of O3) x 816.32653) + 10.20409

SO2 CO NO2 O3

(DDF,1995) Although the IMECA serves its purpose by providing a simple and homogenous measure of the pollution levels in the MCMA, under the current project, only the raw data (the values in their original monitored form and units) are used. These values are more direct and hence provide a higher level of certainty.

90

4.2 Greater Manchester Area (GMA)


The urban area around Manchester is a very complex environment as it is formed by the coalescence of several cities and towns that have their own identity, history and culture. It cannot be properly termed a metropolis because it has more than one city centre; nonetheless, it has the importance and size of one. The area is probably best described as a conurbation, one with a polycentric structure in which ten urban districts interact to create a large urban unit. Its origins lie on a Roman fort established during the first century and spread into one of the first industrial cities of the modern world. Today Manchester or more properly the Greater Manchester Area (GMA) is an effervescent urban agglomeration with many cultural, economic and social challenges and possibilities. The origins of industrial Manchester come from the cotton industry. Its ability to interact as an inland port and with Liverpool, allowed it to develop and dominate the worldwide trade of this product. As this market developed into a mature industry, other supporting industries such as the bleaching and dying of cotton products, chemical manufacturers, engineering goods, and related commerce became established in the region and became a benchmark of the economic importance of the GMA during the current century (Bartholomew, 9 t h ed.). As in the case with the MCMA, the environmental condition of the area can only be assessed when other factors are incorporated into the analysis. Again, the geographic location is paramount and the socio-economic factors and history of the region needs to be considered.

4.21 Geographical Description


The GMA lies in a basin that opens to the west, at an average elevation of 60-m above msl. It is bounded to the east by the Pennine range and the upland spur of Rossendale to the north, hills that rise to over 480-m above msl. This range swings around to the north-west, south-east and, although at much lower altitudes, to the south, approximating the average altitude of the valley. It is situated on the east bank of the River Irwell in south-east Lancashire and its boundaries extend south beyond the River Mersey in northern Cheshire (Wood et al., 1974). It is located in the northern latitudes (5340North, 235W) and has a maritime climate that is affected by a sheltering effect from the Welsh mountains. Consequently, the urban area is less vulnerable to the typical westerly winds associated with its latitudinal location. The usually wet Atlantic air, banked by the Pennine slopes to the east of the area, produces extreme cloudiness throughout the year. Climatologically, the GMA has a lower than average rainfall, especially for a location close to the western seashore (approximately 50 km from Liverpool). The long-term rainfall average stands at a value of 875 mm and is relatively homogeneous throughout the conurbation (Mann, 1994). The temperature is generally without extremes; a January average of 4 C and a summer value of around 15 C (Ravetz, 1997a).

91

Overall, the geographic location of the conurbation contributes to the pollution problems through several of its features, including: Atlantic weather systems bring many anticyclones to the area. These, during winter, translate into low temperatures and hence a low mixed layer over the urban boundary layer. During summer, they promote photolysis of secondary pollutants, resulting mainly in higher O 3 concentrations, Prevailing westerly winds agglomerate air pollutants in the eastern part of the basin, The size of the basin promotes motorised transport within the conurbation. All these, in conjunction with other factors, have made the GMA a relatively polluted urban agglomeration. The large pool of automobiles (estimated to be around 1 million), together with other forms of motorised transport are now the major source of air pollution within the conurbation. It is estimated that the one million cars in the area perform on average six trips each day, amounting to six million car journeys per day. Public transport contributes by performing seven hundred thousand bus trips and seventy thousand local train trips per day (Ravetz, 1997a).

4.22 Factors Affecting Pollution Levels


The history of the development of the GMA is crucial in understanding the pollution problems that affect the area. From the mid 18th to the early 19th century, it established itself as the first industrial city attracting a large base of manufacturing and related commerce that brought with it many environmental pressures. Most of the engines and machines that drove the cotton mills and supporting industry were powered by coal furnaces. These, along with the residential dwellings, steam engines, foundries, dyehouses, etc., emitted large quantities of smoke into the atmosphere of the area. These types of emissions in a west opening basin closed by the Pennines and with the prevailing winds coming from the south-west, had severe consequences for the area and for its population. During the early part of this century, in December 1930, Manchester reported 137 respiratory disease deaths, and in 1931 the cities of Manchester and Salford suffered an intense fog that lasted nine days and resulted in 592 deaths from respiratory disease (Mann, 1994). As the cotton industry declined, many of the mills ceased to operate and were gradually replaced, to a large extent, by heavy and light engineering works that were relatively cleaner and more environmentally friendly, but that brought other problems to the region. Today, as the economic base has changed dramatically, most emissions derive from motorised transport and to a lesser extent the energy generation sector of the economy. In general, it can be stated that the pollution problems in the GMA have closely matched its development as one of the most industrialised centres in Britain, details of these can be found in Williams (1996) and in Mann (1994).

92

The demography of the GMA has changed several times during the last two centuries. The population of the area has gone from immense expansion during periods of economic wellbeing, to contraction at times of austerity. Table 4.7 shows the changes in population, within both Manchester City and the conurbation. As can be seen, there has been a marked flux from the inner part to the surrounding cities and towns, reflecting the change in land use and the deficient infrastructure of the time (Williams, 1996). Table 4.7 Population in Manchester and the GMA Urban Setting Manchester City GMA (in millions) 1801 77,000 0.3 1851 316,000 1.0 1901 544,000 2.2 1951 703,000 2.7 1991 440,000 2.6

(Adapted from Williams, 1996) Also important have been the different reorganisations of its political fabric. From its conception as an industrial centre, the GMA as an urban area has undergone several changes in structure, two in the last twenty years that have infringed in its political organisation and therefore on its dwellings. During 1971, the Royal Commission on Local Government in England (also known as the Redcliffe-Maud Commission) reorganised the political structure of the region and created the Greater Manchester County (GMC). The result was a two-tier system of government consisting of a county level organisation and ten metropolitan districts (Table 4.8), each with different functions and responsibilities. Less than a decade later, the Local Government Act of 1985, enacted in 1986, disintegrated this system and abolished the GMC. The Central Government of England decided to reduce the bureaucratic body of the metropolitan areas and established a system in which the district authorities assumed, individually or collectively, most of the responsibilities and functions of the county level body (Barlow, 1995). Both of these changes are the result of an effort to regenerate the urban area and the economy of the GMA. Table 4.8 Local governments in the GMA, 199 Name Bolton Bury Manchester Oldham Rochdale Salford Stockport Tameside Trafford Wigan Total 93 Population 267,400 176,900 443,600 219,600 210,000 231,000 292,600 216,800 214,700 308,600 2,581,200

4.23 Environmental Controls


Air pollution in the UK can be traced back to the 13th Century when the use of coal was prohibited in London as it was considered to be prejudicial to health. Since this time, regulations have been implemented to improve human health and to promote environmental well-being. In the last two decades, the major sources of legislation targeted towards air quality include several regional, national and Pan-European initiatives. The European Community (EC, and now the European Union, EU), for example, placed limit values on the content of lead (Pb) in petrol fuel (in 1978 the value was set at 0.4 gl-1) and later in 1985 introduced unleaded petrol for the consumption of automobiles (AGMA, 1997). The EC also lead the way in setting limit and guideline values for the most common pollutants, while at the same time harmonising the function of the monitoring networks within the union. At the national level, the Air Quality Regulations of 1989 implemented EC directives pertaining to the limit and guide values for SO2, suspended particles (now PM10), Pb and NOx. The Environmental Protection Act of 1990 (EPA 90) brought small emission sources under air pollution control, and for the first time, delegated control to local authorities and introduced Integrated Pollution Control policies to Britain. In November 1991 the Road Vehicle Regulations instituted vehicle exhaust testing a included it in the annual Ministry of Transport (MOT) tests. Five years later, the Environmental Protection Act of 1995 (EPA 95) required the publication of a National Strategy, which set air quality standards and targets for pollutants of concern, namely: CO, NO2, O3, PM10, SO2, Pb, 1,3 Butadiene and Benzene (AGMA, 1997). At the regional level, the GMA is complying with the national pollution agenda and is investing in programmes intended to reduce the amount of traffic and pollution within the conurbation. The Amalgamation of the Greater Manchester Authorities (AGMA) was created, in part, to undertake inter-district initiatives and to promote local actions to combat environmental problems within the area. With this aim in mind, and under the requirements of the EPA 95, the AGMA developed the Air Quality Management Strategy for Greater Manchester (AGMA, 1997). Included in this strategy are several IPC initiatives that include the expansion of the monitoring network and the inclusion of several of its stations into the DETR national automatic monitoring network, the Automated Urban Network (AUN). Furthermore, the introduction of the light train (Metrolink), connecting some of the suburbs with the city centre in 1992, was intended to reduce the amount of combustion-related transport and therefore the concentration of certain pollutants within the conurbation (HMSO, 1995). Other programs were aimed at tackling area-wide issues, such as urban planning and regeneration. As the city changed and the environmental awareness of its population increased, the legislation governing air quality became more rigorous. 94

Today, the air quality standards are in accordance with the European Union (EU) and are generally very similar to the values recommended by the World Health Organisation (WHO) (MCC, 1997). The standards governing the concentration of the pollutants related to the formation and transformation of O3 and PM10 are presented in Table 4.9 together with their homologous from the WHO. All these pollutants are monitored in four of the seven monitoring stations within a twenty-kilometre radius of the Manchester city-centre. Table 4.9 National Air Quality Strategy, World Health Organisation Standards. Pollutant O3 SO2 NO2 CO National Air Quality Strategy Exceedences Allowed WHO Guidelines * 3% (10 days per year)60 ppb (8hr. mean) 0.1% (8 hours per year)175 ppb (10-min) 44 ppb (24-hr) 17 ppb (annual) 150 ppb (1-hour mean) 21 ppb (annual mean) 110 ppb (1-hr) 21-26 ppb (annual) 10 ppm (8-hour mean) 90 ppm (15-min) 50 ppm (30-min) 25 ppm (1-hr.) 10 ppm (8-hr) 50 g / m3 (24-hour mean) 1% (88 hours per year) Air Quality Standard 50 ppb (8-hour mean) 100 ppb (15 minute mean) (Adaptec from MCC, 1997)

PM10

* World Health Organisation

4.24 Automated Urban Network of the GMA


Air pollution monitoring in Manchester began during 1959, measuringthe concentration of smoke and sulphur dioxide (SO2). Since then, the monitoring of pollutants has expanded to cover most of the GMA, including one meteorological station in the south of the region, and to form part of the Automatic Monitoring Network (AUN) of the United Kingdom (MCC, 1997). The national network began operation in 1992 with six stations and currently incorporates more than 89 sites, wholly or partially funded by the Department of the Environment Transport and Regions (DETR). The whole of the AUN is divided into five different units that are operated by different organisations on behalf of the DETR. The Central Management and Co-ordinating Unit (CMCU) is run by TBV Science and is responsible for receiving the data by telemetry from each site and for monitoring their output and carrying out diagnostic test for signs of malfunction within the network.

95

The CMCU passes all the information to the Data Dissemination Unit (DDU) at the National Environmental Technology Centre (NETCEN) and this relays the information to the public either via CEEFAX/Tele-text or through the i n t e r n e t ( h t t p : / / w w w. a e a t . c o . u k / n e t c e n / a i r q u a l / i n d e x . h t m l ) . Detailed calibration of all the monitoring instruments and site practices are audited by the Quality Assurance and Quality Control Unit (QA/QC), undertaken by the National Physical Laboratory (NPL) on a quarterly basis. They are responsible for maximising the accuracy and comparability of the data and for producing the data sets in their final published form. Other key positions within the network include the Local Site Operator (LSO), who visits each site at least once a month to carry out routine instrument maintenance and checks, and the Equipment Support Unit (ESU), responsible for servicing the equipment and for repairing or replacing instruments as necessary (PORG, 1997). The quality of the data is ensured through the instrument calibration and maintenance program and through the data inspection and review practices. Daily automatic calibration and checks on the analysers are monitored by the CMCU, while every month a more thorough check is made and reported by the LSO. The analysers are serviced every six months by the ESU and before and after each of these services, the QA/QC unit visits each site with a primary standard to calibrate each analyser. The primary standards are compared with other national and international standards to assure the quality and precision of the data. The data disseminated rapidly by the DDU has a provisional zero offset applied and is checked for anomalous data points automatically and manually. An accurate estimate of the zero offset and the overall behaviour of the output from the instruments require information from subsequent calibrations, most of which are carried out every six months by the QA/QC unit. Information form the automatic calibrations, the LSO visits, service records and any other relevant information are examined and the calculated zero offset and other scaling factors are applied to the data, where appropriate. When significant uncertainties associated with the data or the concentrations measured by the monitoring equipment are present, the data are rejected. Before the provisionally ratified data is published, it is scrutinised along with any ancillary information by a quality circle including senior project staff of the QA/QC unit (PORG, 1997).

4.241 Monitoring Stations in the GMA


Within the GMA there are several monitoring stations, seven of which belong to the AUN and monitor most of the atmospheric pollutants necessary for the current project. A description of the sites, along with a list of the pollutants that they measure, is presented in Table 4.10.

96

Table 4.10 GMA Monitoring Stations Site Name Bolton Bury Piccadilly Salford Eccles Stockport South Town Hall Altitude (above msl)
103 m 50 m 45 m 30 m 62 m 68 m 90 m

Pollutants Measured
O3, NOx, NO2, NO, CO, SO2, PM10 O3, NOx, NO2, NO, CO, SO 2 , PM 10 O3, NOx, NO2, NO, CO, SO2, PM10 O3, NOx, NO2, NO, CO, SO 2 , PM 10 NOx, NO2, NO, CO, SO2, PM10 O3, NOx, NO2, NO, SO2, Meteorology NOx, NO2, NO, CO

Site Description
Residential area, 110 m from main road. Located 30 m from M62 motorway (200000 v/d)* Busy city centre, 20 m. from main road Parkland immediately to the south, 7 m from road. On roof of 2 storey building on one way St. Within airport boundary, 1 km from runway Elevated city centre location, 20 m from road.

Type of Site
Urban Background Roadside Urban Centre Urban Industrial Urban Background Suburban Urban Background

*Vehicles per day The location of the stations can be seen in Figure 7.1. In general, the monitoring equipment used in the AUN stations is the same that is used in the MCMA (ultraviolet light photometry for O3, non-dispersive infrared spectroscopy for CO, pulsating fluorescence for SO2, chemiluminescence for NOx (including NO, and NO2 by subtraction) and attenuation via beta radiation for PM10). The same is true for the meteorological instruments, a conventional anemometer and wind vane are used to measure wind velocity and direction, while the temperature is measured with a thermistor, and the relative humidity by hygrometer (http//www.meto.gov.uk/sec2/pg3/awsmt.html, February, 1999).

4.3 Conclusion
Air quality in urban agglomerations is affected by an immensenumber of factors. Among some of the most important are the location, climate, and meteorology of the area. Socio-economic, demographic and historical parameters should also be included in any environmental analysis, as they tend to either ameliorate or aggravate the environmental health of the region. Particular to the MCMA are its location, high population density, large socio-economic differentials and relative economic importance in relation to the rest of the country. These have translated into an urban area that has severe pollution problems, especially in relation to its large industrial and commercial base, the large pool of automobiles and the overcrowded infrastructure.

97

Together, and coupled with its geographical location, in a high altitude closed valley in the middle of a country that is bounded by two oceans, they exacerbate the pollutant levels of the capital city. The GMA on the other hand, lies in a basin that opens to the west, against the prevailing wind. The conurbation has historically been polluted by its industrial base, but now the problem has shifted to pollution arising from the transport sector. The size and spatial organisation of the urban area promotes the use of private transport and it has therefore created a large pool of automobiles that operate and pollute on a daily basis. As the basin is bounded by the Pennine range to the west, and the location is affected by the usually wet Atlantic air, anticyclonic conditions promoting atmospheric stability, are common throughout the area. In both agglomerations, pollution problems have lead each government to regulate emissions and promote initiatives to control the ambient concentration of the most common pollutants. This has translated into the establishment of different strategies aimed at reducing the environmental hazards inherent to large urban agglomerations. Specifically, both areas have programs to reduce photochemical pollution (using catalytic converters in cars), to phase out leaded petrol and to substantially reduce the sulphur content in diesel fuels. At the same time, large amounts of resources have been allocated to quantify the concentration of atmospheric pollutants, via monitoring networks and research projects, and to diminish the negative effects of atmospheric pollutants. These efforts have translated into a better understanding of the problems that affect each area and various strategies are currently being developed to ameliorate them.

98

Results from the MCMA


Chapter Five

5 Introduction
Annual time series for the concentration of PM10, O3, their precursors and meteorological parameters were calculated in monthly hourly periods between 1994 and 1996 in five selected stations of the monitoring network of the MCMA (RAMA). The stations are located between the cardinal points of the city and one in the city centre (see Figure 5.1). Figure 5.1 Map of the MCMA and the location of the Stations used in the study
N

Mexico City Monitoring Network (RAMA) Geographic Distribution of Stations


Mexico City Metropolitan Area Municipal Division IMECA Zoning Monitoring Station

NW
TLA

NE
XAL

MER

CC
PED CES

SW

SE

Scale
0 5 10 Km

Annual and tri-annual time series for PM10, sulphur dioxide (SO2), ozone (O3), nitrogen dioxide (NO2), and nitric oxide (NO) were graphed in order to understand the associations between them. Meteorological parameters were incorporated into the analysis to illustrate their influence on the concentration of PM10 and O3. Diurnal distributions of these same pollutants were plotted to study the different levels of activity in the city and the influence of solar radiation on the production of primary and secondary pollutants. Furthermore, to try to comprehend some of the formation mechanisms and the associations between the pollutants named above, statistical analyses were performed to the available data. The concentration and the frequency of wind patterns were analysed by plotting rose diagrams of pollutant vs. wind direction and wind frequency vs. wind direction. All the work mentioned above has the aim of understanding the sources, associations and formation processes of PM10 and O3 in the MCMA. In the following pages, all the analyses will be placed into context and they will be discussed in the following chapter. 101

5. 1 Time Series
The monthly hourly average concentration time series represent the mean concentration of the hourly averages each month in all the stations. They show, among other things, the distribution of the concentrations throughout the year and between the years analysed.

5.11 PM10 Time Series


The PM10 time series for the period 1994 to 1996 is displayed in Figure 5.2. It shows among other things, an increase in concentration from one year to the next. Figure 5.3 displays the yearly averages and as can be seen in the plot, the trends followed sharp seasonal fluctuations.

Figure 5.2 PM10 Time Series 1994-1996

1994-96 1994-96 (excluding XAL)

100 90 80 70 60

PM10

50 40 30 20 10 0 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

Months

102

Figure 5.3 PM10 Yearly Trend excluding XAL

1994 (Excluding XAL, 55.43 ug/m3 1994 (Excluding XAL, 46.60 ug/m3 1994 (Excluding XAL, 61. 75 ug/m3

90 80 70 60

ug/m3

50 40 30 20 10 0 Jan Feb Mar Apr May Jun Jul Ago Sep Oct Nov Dec

Figure 5.4 PM10 Monthly Averages, 1994

120

100

TLA XAL MER PED CES

80

ug/m3

60

40

20

0 Jan Feb Mar Apr May Jun Jul Ago Sep Oct Nov Dec

103

Figure 5.5 PM10 Monthly Averages 1995


TLA XAL MER PED CES

160 140 120 100

ug/m3

80 60 40 20 0 Jan Feb Mar Apr May Jun Jul Ago Sep Oct Nov Dec

Figure 5.6 PM10 Monthly Averages 1996

250

200

TLA XAL MER PED CES

150

ug/m3

100

50

0 Jan Feb Mar Apr May Jun Jul Ago Sep Oct Nov Dec

104

In these figures, one station can be distinguished from the rest as it completely breaks its normal pattern. Xalostoc, the station in the north east of the city, reported some of the highest average concentrations of PM10 especially during the autumn winter periods of 1995 and 1996. In the autumn-winter period of 1994 however, the concentration dropped to the lowest of all the stations. The relatively low values reported during the late winter period of 1994 can be partially attributed to meteorological parameters, but it is also believed that this station was affected by a very localised source making it vulnerable to sporadic changes. This belief is based on the location of the station; it is situated inside a Volkswagen dealership next to the engine cleaning section (Figure 5.7). The relationship between PM10 and wind speed also supports this hypothesis, as it is the only station that reported a strong negative relationship with increasing wind speed. The rest either did not follow it or the relationship was very weak. Table 5.1 shows the Coefficients of Determination (COD, R2) for the respective monitoring stations. Figure 5.7 Photograph of the Xalostoc Station in the Northeast

Table 5.1 PM10 vs. Wind Speed, 1994

Station
Tlanepantla (TLA) Xalostoc (XAL) Merced (MER) Pedregal (PED) Cerro de la Estrella (CES)

Coefficient of Determination (R2) 0.032 0.561 0.002 0.066 0.104

Figure 5.8 shows the relationship between PM 10 and wind speed.

105

Table 5.8 PM10 vs. Wind Speed, 1994


PM10 1994-96 (excluding XAL) WSP 1994-96

90 80

5 70 60 4

ug/m3

50 3 40 30 20 1 10 0
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

m/s

Months

In this figure, it can be seen that the PM10 concentration was influenced by wind speed only during 1994, while in the two other years the relationship was not obvious. Figure 5.9 shows the relationship between relative humidity, temperature and the concentration of PM10 during 1994, 1995 and 1996. Figure 5.9 PM10, Temperature and Relative Humidity, 1994-96
PM10 1994-96 RH 1994-96 Temp 1994-96

90 80 70 60

25

20

15

ug/m3

50 40 10 30 20 10 0
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

Deg. C

Months

106

5.12 Ozone Time Series


Ozone is a secondary pollutant mainly produced by the photolysis of NO2, the details of which are explained in chapter three. Ozone has been categorised as the most severe pollution problem of the MCMA both because of its high concentrations and because of the frequency of departures from the MAQS. In the years analysed the norm (0.11 ppm = 100 IMECA points) was exceeded 95% of the days in 1994, 89% in 1995 and 89% in 1996 (INE, 1997). Figure 5.10 shows the three-year time series for O3 during 1994-96 and Figure 5.11 depicts the yearly distribution during the same period.

Figure 5.10 O3 Time Series 1994-1996

1994-96

0.07 0.06 0.05 0.04

ppm

0.03 0.02 0.01 0


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

Months

107

Figure 5.11 O3 Yearly Trend


1994 (0.045 ppm) 1994 (0.043 ppm) 1994 (0.036 ppm)

0.070 0.060 0.050 0.040

ppm
0.030 0.020 0.010 0.000 Jan Feb Mar Apr May Jun Jul Ago Sep Oct Nov Dec

O3 series have a definite seasonal variation but it is related to the solar radiation budget and the availability of NO2 and organics in the troposphere. Figures 5.12 to 5.14 show the monthly averages by station in each of the years analysed. Figure 5.12 O3 Monthly Average 1994

0.070

0.060

TLA XAL MER PED CES

0.050

0.040

ppm

0.030

0.020

0.010

0.000 Jan Feb Mar Apr May Jun Jul Ago Sep Oct Nov Dec

108

Figure 5.13 O3 Monthly Average 1995


0.080 0.070 0.060 0.050

TLA XAL MER PED CES

ppm

0.040 0.030 0.020 0.010 0.000 Jan Feb Mar Apr May Jun Jul Ago Sep Oct Nov Dec

Figure 5.14 O3 Monthly Average 1996

0.070

0.060

TLA XAL MER PED CES

0.050

0.040

ppm

0.030

0.020

0.010

0.000 Jan Feb Mar Apr May Jun Jul Ago Sep Oct Nov Dec

The station that persistently reports the highest values is Pedregal (PED) in the Southwest of the MCMA. This station is located near the National University (UNAM), away from strongest NO2 sources.

109

It is, therefore, likely that the prevailing wind direction of the city (north to south) and the meteorology are triggering the high values in this station.

5.13 Precursor Species


Secondary particles and O3 require the presence of precursor species to trigger their formation in the atmosphere. Specifically, PM10 may form when SO2 or NO2 oxidise to form H2SO4 and HNO3 respectively, while O3 may only form through the photolysis of NO2. NO, on the other hand, is important because it titrates O3, providing an important removal mechanism. In the following section, the tri-annual and yearly trends of these three pollutants will be presented. Sulphur dioxide is a primary pollutant and a major source of secondary aerosols. Figure 5.15 shows the tri-annual time series for SO2 for the period between 1994 and 1996. It can be observed that the overall monthly-hourly average concentration is generally decreasing. Figure 5.16 displays the yearly average during the same period. Figure 5.15 SO2 Time Series 1994-1996
1994-96

0.030

0.025

0.020

ppm

0.015

0.010

0.005

0.000
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

Months

110

Figure 5.16 SO2 Yearly Trends

1994 (0.020 ppm) 1994 (0.015 ppm) 1994 (0.016 ppm)

0.030

0.025

0.020

ppm

0.015

0.010

0.005

0.000 Jan Feb Mar Apr May Jun Jul Ago Sep Oct Nov Dec

In the above figure, it can be observed that the concentrations of 1995 and 1996 were lower than in 1994 and that in half of the months, 1996 was lower than 1995. No seasonal variation is apparent except for a slight decrease during the summer, especially during 1995.Nitrogen Oxides (NOx) is the term used to refer to the combination of nitric oxide (NO) and nitrogen dioxide (NO2). The monitoring network of the MCMA measures NOx and NO while NO2 is obtained by subtraction of the two former pollutants (NO2 = NOx NO). It is important, however, to analyse NOx individually as NO and NO2 are responsible for a large proportion of the production and removal of secondary pollutants, i.e. O3 and aerosols. The principal precursor of O3 and of secondary aerosols is NO2. Figure 5.17 shows the NO2 time series for the 1994-1996 period and Figure 5.18 displays the yearly averages. As before, all the trends represent the mean concentration of the hourly averages. It can be noted that, in both figures, the seasonal variation is very strong. The concentration began high in the early winter, descended during the spring, reached a minimum in the summer and then ascended during the autumn and winter periods. This trend is very similar to the PM10 time series. Although NO2 is the principal precursor of both O3 and secondary PM10, its existence in the atmosphere is mainly due to the oxidation of NO. The nitric oxide trend during 1994-96 is shown in Figure 5.19, while the three-year averages are displayed in Figure 5.20.

111

Figure 5.17 NO2 Time Series 1994-1996

1994-96

0.030

0.025

0.020

ppm

0.015

0.010

0.005

0.000
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

Months

Figure 5.18 NO2 Yearly Trend

1994 (0.020 ppm) 1994 (0.015 ppm) 1994 (0.016 ppm)

0.070

0.060

0.050

0.040

ppm
0.030 0.020 0.010 0.000 Jan Feb Mar Apr May Jun Jul Ago Sep Oct Nov Dec

112

Figure 5.19 NO Time Series 1994-1996


1994-96

0.030

0.025

0.020

ppm

0.015

0.010

0.005

0.000
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

Months

Figure 5.20 NO Yearly Trend


1994 (0.020 ppm) 1994 (0.015 ppm) 1994 (0.016 ppm)

0.080 0.070 0.060 0.050

ppm

0.040 0.030 0.020 0.010 0.000 Jan Feb Mar Apr May Jun Jul Ago Sep Oct Nov Dec

The relationship between the particles and SO2 and NO2 is important, as the secondary particles are mainly formed through gaseous and aqueous phase oxidation of these two species. Figure 5.21 shows their tri-annual time series and it can be observed that PM10 is not related to SO2, but the trend of NO2 is relatively similar to the particulate time series.

113

They both have marked seasonal fluctuations. Furthermore, the relationship between O3, NO2 and NO is important as these drive its concentration in the atmosphere.As explained before, NO2 photolyses to produce O3, while NO titrates and removes it from the troposphere. Figure 5.22 displays their relationship. Figure 5.21 PM10, SO2 and NO2 Time Series, 1994-96
PM10 1994-96 SO2 1994-96 NO2 1994-96

100 90 80 70 60

0.07 0.06 0.05 0.04 0.03 0.02 0.01 0 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

ug/m3

ppm

50 40 30 20 10 0

Months

Figure 5.22 O3, NO2 and NO, 1994-96


O3 1994-96 NO2 1994-96 NO 1994-96

0.08 0.07 0.06 0.05

ppm

0.04 0.03 0.02 0.01 0


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

Months

114

5.15 Meteorology Time Series


The concentration of pollutants in the atmosphere is highly dependent on the meteorological conditions of the troposphere. The meteorological parameters analysed here are temperature, relative humidity, rainfall, and wind speed. The temperature time series for the MCMA in the years analysed is presented in Figure 5.23, while the yearly profiles are displayed in Figure 5.24. Figure 5.23 Temperature Time Series 1994-1996
1994-96

25

20

Degrees Celsius

15

10

0
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

Months

Figure 5.24 Temperature Yearly Trend


1994 (16.63) 1995 (16.55) 1996 (16.25)

25

20

Degrees Celsius

15

10

0 Jan Feb Mar Apr May Jun Jul Ago Sep Oct Nov Dec

115

Relative humidity (RH) is a measurement of the amount of moisture in the atmosphere and as such, it influences several pollutant formation and removal mechanisms. Figure 5.25 shows the relative humidity time series from 1994 to 1996 and Figure 5.26 displays the yearly averages. Figure 5.25 RH Time Series 1994-1996
1994-96

90 80 70 60

Percentage

50 40 30 20 10 0
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

Months

Figure 5.26 RH Yearly Trend


1994 (60.92%) 1995 (54.58%) 1996 (49.96%)

90 80 70 60

Percentage

50 40 30 20 10 0 Jan Feb Mar Apr May Jun Jul Ago Sep Oct Nov Dec

116

Wind is a result of many processes involved in the atmosphere and it enhances dispersion by inducing mechanical turbulence. High wind speeds are therefore associated with better dispersion and in general lower loadings of primary pollutants. The wind speed time series for the period under study is presented in Figure 5.27 and the yearly trends in Figure 5.28. In the MCMA the average wind speed ranges from 1-2 m/s, but in 1994 the oceanic effect of "el nio" enhanced this average to reach a range of 3-5 m/s. Figure 5.27 Wind Speed Time Series 1994-1996
1994-96

m/sec.

0
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

Months

Figure 5.28 Wind Speed Yearly Average


1994 (4.10m/sec.) 1995 (2.17 m/sec.) 1996 (1.49 m/sec.)

m/sec.

0 Jan Feb Mar Apr May Jun Jul Ago Sep Oct Nov Dec

117

As it is obvious from the two Figures, there is a big variation on wind speed between 1994, part of 1995 and 1996. The variation, as stated, is attributed to El nio a cyclic effect that changes the synoptic or macro-scale meteorological patterns around the globe every three to seven years. As part of the time series analyses, precipitation data needs to be incorporated. The yearly statistics are given in Table 5.2 and is evident from the values that the total precipitation throughout the year was much lower during 1996 than in 1994 and 1995, even after the last two months are subtracted from the first two years. Also important is the distribution: most of the rain fell during the summer, while the winter and equinoxes were relatively dry. Figures 5.29 shows the yearly trends of pluvial precipitation during 199496, while Figure 5.30 illustrates the monthly distribution of rain during the same years. Finally, Figure 5.31 depicts the cumulative frequency of rain during the period under study.

Table 5.2 Rainfall Data, 1994-96 1994 11.03 1.47 2.58 32.40 52.02 156.21 146.84 193.73 162.44 71.92 5.66 4.52 840.82 1995 24.42 11.54 11.04 8.27 66.76 159.25 134.79 198.56 107.19 38.55 47.77 53.33 861.47 1996 0.00 0.26 2.18 26.52 20.75 113.10 115.50 114.65 129.23 41.91 N/A N/A 564.10 Ave. 1994 - 96 11.82 4.42 5.27 22.40 46.51 142.85 132.38 168.98 132.95 50.79 26.72 28.93

Jan. Feb. March April May June July Aug. Sep. Oct. Nov. Dec. Total:

118

Figure 5.29 Precipitation Time Series, 1994-96

Ave. 1994.96

250

200

mm of rain

150

100

50

0
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

Months

Figure 5.30 Monthly Precipitation 1994-96

1994 1995 1996 Ave. 1994-96

250

200

mm of rain

150

100

50

0 Jan Feb Mar Apr May Jun Jul Ago Sep Oct Nov Dec No data for Nov. and Dec., 1996

119

Figure 5.31 Cumulative Frequency, 1994-96


1994 1995 1996 Ave. 1994-96

900 800 700 600

mm

500 400 300 200 100 0 Jan Feb Mar Apr May Jun Jul Ago Sep Oct Nov Dec

5.2 Diurnal Analyses


Diurnal profiles for the concentration of the main pollutants under study were calculated to evaluate the influence social of activity in the city and of the solar radiation on the concentration of the contaminants. A large proportion of the pollutants emitted in the MCMA are from automobile sources; it is estimated that 75% of all emissions are from the transport sector (INE, 97), therefore the level of social activity in the city is critical to the analysis of formation mechanisms (see Table 5.3). On the other hand, photolytic contaminants need solar radiation to start the reaction processes and this will only occur during the sunlight hours of the day. Table 5.3 Emission Inventory of the MCMA by Sector (Quantities in Thousands per annum) Category Industry Services Soil and Vegetation Total 452 100 46 100 2,358 100 129 100 1,026 100 4,010 100 PM10 Ton % Ton 6.3 1.4 1.1 0.2 425.3 94.2 SO2 % 26.1 57.3 7.2 15.9 CO Ton 8.7 0.9 % 0.1 NOx Ton 5.3 % HC Ton % Total Ton 413.0 464.2 % 3 10 75 12 3.2 105.7

0.4 31.5 24.5 33.1

4.2 398.4 38.9 38.9 3.8

Transport 18.8 4.2

12.2 26.8 2348.5 99.5 91.8 71.3 555.3 54.1 3026.6

120

Diurnal trends for all the pollutants under study were calculated during the day-time (07:00 20:00 hrs.) and night-time hours (21:00 06:00 hrs.) taking into consideration not only the hours of sunlight, but also the peak hours of social activity. Figures 5.32 to 5.36 show the diurnal trends for PM10, SO2, O3, NO2 and NO, respectively. Figure 5.32 PM10 Diurnal Variation
7-20 hrs. 21-06 hrs.

120

100

80

ug/m3

60

40

20

0
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

Months

Figure 5.33 SO2 Diurnal Variation


7-20 hrs. 21-06 hrs.

0.030

0.025

0.020

ppm

0.015

0.010

0.005

0.000
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

Months

121

Figure 5.34 O3 Diurnal Variation


7-20 hrs. 21-06 hrs.

00.90 0.080 0.070 0.060 0.050 0.040 0.030 0.020 0.010 0.000
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

ppm

Months

Figure 5.35 NO2 Diurnal Variation


7-20 hrs. 21-06 hrs.

0.080 0.070 0.060 0.050

ppm

0.040 0.030 0.020 0.010 0.000


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

Months

122

Figure 5.36 NO Diurnal Variation


7-20 hrs. 21-06 hrs.

00.90 0.080 0.070 0.060

ppm

0.050 0.040 0.030 0.020 0.010 0.000


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

Months

5.3 Statistical Analysis


Statistical analyses were performed to the data of 1996 in order to understand the relationship among the pollutants and the influence of meteorological parameters on these. Ordinary Least Squares Regression (OLS) analysis was used to evaluate the Coefficient of Determination (COD or R2) between PM10 and SO2, NO2, temperature, and relative humidity. Multiple regression was then applied to the parameters with a relatively high correlation coefficient in order to evaluate the relationship together. The same type of analyses were performed between O3 and NO2, NO, temperature and relative humidity and again multiple regression was applied to the relevant parameters. Table 5.4 shows the CODs for the relationships and for the multiple regression analyses.

123

Table 5.4 Statistical Analyses and Durbin Watson Test

Pollutant SO2 NO2 Temp Rel. Hum. (RH) Mult. Reg. (NO2 and RH) NO2 NO Temp Rel. Hum. (RH) Mult. Reg. (NO2, and RH)

COD (R2) 0.059 0.31 0.002 0.31 0.44 0.23 0.0002 0.084 0.20 0.49

Significance (Statistic) P = .99 (23.13) P = .99 (165.50) P = .63 (0.81) P = .99 (162.70) P = .99 (95.84) P = .99 (114.81) P = .20(0.06) P = .99 (33.32) P = .99 (91.06) P = .99 (117.10)

Durbin-Watson Test. 0.425 0.853 0.446 0.726 0.652 0.905 0.853 0.885 1.050 1.007

PM10

O3

As it can be observed, the relationship between some of the pollutants is relatively strong, especially between the NO2, relative humidity and both PM10 and O3. The analysis however, may be flawed as the Durbin Watson test stipulates that the residuals in the data autocorrelated, in other words all the estimators of the regression may be invalid including the t and F statistics (Clark and Husking, 1986). On the other hand, the relationships are correct and although they cannot be tested for significance, they show the strength of the association between PM10, O3, their precursor species and some of the meteorological parameters. Moreover, the relationship plots can be used to determine the formation conditions of both PM10 and O3. Figures 5.37 to 5.40 show the relationship between PM10 and SO2, NO2, temperature, and relative humidity respectively.

124

Figure 5.37 PM10 vs. SO2


SO2 y= 1027.2x + 56.998 OLS R 2 = 0.0598
160 140 120 100
PM10

80 60 40 20 0 0.000 0.010 0.020 0.030 0.040 0.050 0.060 0.070 0.080 0.090

Figure 5.38 PM10 vs. NO2

NO2 y= 1028.2x + 24.811 OLS R 2 = 0.3126


160 140 120 100
PM10

80 60 40 20 0 0.000

0.010

0.020

0.030

0.040
NO2

0.050

0.060

0.070

0.080

0.090

125

Figure 5.39 PM10 vs. Temperature


Temp OLS
160 140 120 100
PM10

y= 0.5324x + 65.288 R 2 = 0.0022

80 60 40 20 0 0 5 10 15 20 25

Temp.

Figure 5.40 PM10 vs. Relative Humidity


RH OLS
160 140 120 100 80 60 40 20 0 0 10 20 30 40 50 60 70 80 90

y= 1.1271x + 130.21 R 2 = 0.3089

PM10

RH

126

Figure 5.41 to 5.44 show the relationships between O3 and NO2, NO, temperature and relative humidity. Figure 5.41 O3 vs. NO2, 1996
NO2 OLS
0.07

y= 0.3862x + 0.0204 R 2 = 0.2398

0.06

0.05

0.04

O3

0.03

0.02

0.01

0.00 0.000 0.010 0.020 0.030 0.040 0.050 0.060 0.070 0.080 0.090

NO2

Figure 5.42 O3 vs. NO, 1996


NO OLS y= 0.0064x + 0.0358 R 2 = 0.0002

0.07

0.06

0.05

0.04

O3

0.03

0.02

0.01

0.00 0.000 0.010 0.020 0.030 0.040 0.050 0.060 0.070 0.080 0.090

NO2

127

Figure 5.43 O3 vs. Temperature


Temp OLS
0.07

y= 0.0012x + 0.0167 R 2 = 0.0839

0.06

0.05

0.04

O3
0.03 0.02 0.01 0.00 0 5 10 15 20 25

Temp

Figure 5.44 O3 vs. Relative Humidity


RH OLS
0.07

y= 0.0003x + 0.0526 R 2 = 0.2001

0.06

0.05

0.04

O3
0.03 0.02 0.01 0.00 0 10 20 30 40 50 60 70 80 90

RH

128

5.4 Wind Analyses


Pollutant roses were prepared to examine the sources of PM10, and O3 together with the species affecting their concentration, namely SO2, NO and NO2. Roses for each station were calculated by separating the concentration of each of the above contaminants into eight categories, related to the cardinal points. The categories or bins were: north (338 22), north-east (23 67), east (68 112), south-east (113 157), south (158 202), south-west (203 247), west (248 292), and north-west (293 337). In the prevailing wind roses, the frequency of occurrence was graphed against the bins mentioned above, while for the pollution roses the average concentration in each was used. Figure 5.45 shows the prevailing wind direction in the MCMA during 1994, 1995 and 1996, while Figure 5.46 displays the aggregated data for the same period.

Figure 5.45 Prevailing Wind Direction 1994-96

Wind Frequency 1994 Wind Frequency 1995 Wind Frequency 1996

129

Figure 5.46 Wind Frequency, 1994-96

N 35000 30000 NW 25000 20000 15000 10000 5000 W 0 E NE

Wind Frequency 1994-96

SW

SE

These graphs clearly show that the wind is mainly coming from the orthern part of the city. Pollutant roses were prepared for PM10 and its precursors during the same period (1994 1996) for each of the stations and are shown in Figures 5.47 to 5.49.

Figure 5.47 PM10 vs. Wind Direction, 1994-96

TLA PM10, 1994 - 96 XAL Pm10, 1994 - 96 MER PM10, 1994 - 96 PED PM10, 1994 - 96 CES PM10, 1994 - 96

130

Figure 5.48 SO2 vs. Wind Direction, 1994-96

TLA SO2, 1994 - 96 XAL SO2, 1994 - 96 MER SO2, 1994 - 96 PED SO2, 1994 - 96 CES SO2, 1994 - 96

Figure 5.49 NO2 vs. Wind Direction, 1994-96

TLA NO2, 1994 - 96 XAL NO2, 1994 - 96 MER NO2, 1994 - 96 PED NO2, 1994 - 96 CES NO2, 1994 - 96

Ozone and nitric oxide pollutant roses are shown in Figures 5.50 and 5.51.

131

Figure 5.50 O3 vs. Wind Direction, 1994-96

TLA O3, 1994 - 96 XAL O3, 1994 - 96 MER O3, 1994 - 96 PED O3, 1994 - 96 CES O3, 1994 - 96

Figure 5.51 NO vs. Wind Direction, 1994-96


N 0.09 0.08 NW 0.07 0.06 0.05 0.04 0.03 0.02 0.01 0 NE

TLA NO, 1994 - 96 XAL NO, 1994 - 96 MER NO, 1994 - 96 PED NO, 1994 - 96 CES NO, 1994 - 96

SW

SE

5.5 Conclusion
All the work presented above has the purpose of analysing atmospheric pollution in the MCMA, especially the formation and transformation mechanisms of PM10 and O3. Time series and diurnal distributions will be used to interpret the behaviour of the pool of pollutants examined, while the statistical analyses and the wind roses will help to gain an understanding of the associations and their sources. In the next chapter all the work will be placed into perspective and the associations, behaviour, sources and formation mechanisms of the particles and ozone in the MCMA will be examined. 132

Discussion of the Results from the MCMA


Chapter Six

6 Introduction
The results presented in the previous chapter will be discussed in the following sections. The time series of PM10 and O3 will be related to meteorological parameters to understand the influence that temperature, relative humidity, pluvial precipitation and wind speed have on the overall loadings of these two pollutants. The associations that PM10 and O3 have with their precursor species will also be examined. NO2 and SO2 will be associated to secondary particles, while O3 will be related to NO and NO2. The influence of social activity and the level of insolation will be studied through diurnal variation charts. These plots display the concentration of the pollutants involved throughout the day and the night. A quantitative estimate of the relationships between the particles, ozone, their precursors and meteorological parameters will be attempted through regression analyses. Finally, the influence of wind direction will be examined through pollutant and wind roses. All the information will be placed into context and will be related to the formation and transformation of PM10 and O3 within the MCMA.

6.1 Time Series


Time series for the period of 1994 to 1996 were calculated using the data from five monitoring stations within the MCMA, the locations of which are depicted in Figure 5.1. The series presented in the preceding chapter constitute triannual and annual trends of PM10, SO2, O3, NO2, NO and meteorological parameters, all of which will be discussed in the following sections.

6.11 PM10
The station in the north-east of the city, Xalostoc (XAL) persistently reported some of the highest particle loadings in the MCMA. This station is believed to be affected by a very localised source, thus being vulnerable to sudden local emission changes. As it was stated in the previous chapter, this station is situated inside a Volkswagen dealership, next to the engine cleaning section of the garage (Figure 5.7), thus high particle loadings are emitted sporadically. Table 5.1 provides the Coefficients of Determination (COD or R2) for the relationship between PM10 and wind speed in the five different stations. The only station with a relatively high COD (R2 = 0.56) was XAL, indicating that the sources were close to the monitoring equipment (de Icaza, Choularton, 1998). Figure 5.4 displays the particle loadings per station during 1994 and it can be observed that the trends in all the stations are very dissimilar, indicating that the processes affecting their concentrations are different. Figures 5.5 and 5.6 show a similar scenario during 1995 and 1996 but in these, the XAL station can be distinguished because it persistently reported some of the highest concentrations in the area.

135

The concentrations reported in XAL ranged from 57 to 196 g/m3 during the last two years, while in the rest of the stations the lowest loading was 21 g/m3 and the highest was 135 g/m3. Figure 5.2 displays the tri-annual trend of PM10 and it includes the curve representing the mean particle concentration excluding the XAL station. As can be observed, the average loadings during 1994 were relatively similar, while in the latter two years the differences were much greater. Because of the unreliability of the measurements, the values reported in the XAL station will be excluded from the analysis. The tri-annual trend of PM10 shows a highly seasonal distribution, higher in the winter and lower in the summer months, shown in Figure 5.2. In general, the concentration began high each winter period: January to March with averages ranging from 33 to 75 g/m3. The PM10 loadings descended during the spring equinox to reach its lowest values during the summer: May to August with concentrations ranging from 34 to 73 g/m3. After this, the concentrations began to increase in the autumn and reached a peak again during the early part of the winter: November to December with averages ranging from 44 to 64 g / m3. This trend was followed in 1995 and 1996, but in 1994, the concentrations during September did not increase (de Icaza, Choularton, 1998). The reason explaining this change is centred on the meteorology of the period. In this particular month, the amount of precipitation was higher than the average, registering a total of 162 mm of rain, while during 1995, and 1997 the amount for September totalled 107 and 129 mm, as shown in Table 5.2. Further reasons can be found in the wind speed time series. In the first year analysed and the first quarter of 1995 wind speed was in the order of 3 to 5 meters per second (m/s), while in the rest of the period the mean wind speed ranged from 1 to 2 m/s, shown in Figure 5.27. The explanation for this dramatic reduction is that during 1994 the natural phenomenon of el nio was present. This event is a reversal of weather patterns in the Southern Hemisphere that typically enhances ocean temperatures (2 8 C above normal) and is usually accompanied by altered weather patterns around the globe. In North America and Mexico in particular, it has the effect of either producing excessive rain or droughts, and it can induce unusual winter weather. El nio events recur every three to seven years at irregular intervals. El nio may occur any year and it is not predictable, although this situation is unusual. During 1994, the weather anomalies attributed to this phenomenon were mild creating only non-catastrophic changes. In the MCMA, this translated into unusual wind speeds, which in turn affected the whole atmosphere of the city. In the yearly trend (Figure 5.3), it can be observed that during 1994 the seasonal variation was smaller and that the mean concentration was lower, probably related to the stronger winds causing more ventilation and greater mixing depths during 1994 (de Icaza, Choularton, 1998).

136

The effect that higher wind speeds had on the distribution of the particles indicates that there is a negative relationship between PM10 concentration and wind speed. This relationship, however, is only valid during 1994, indicating that there is a threshold value at which the relationship stands. In the higher regime, when wind speed was in the order of 3m/s 5m/s, increasing wind speed translated into lower particle loadings, as shown in Figure 5.8. When the wind speed was lower, in the order of 1m/s 2 m/s, there was no apparent relationship. Table 6.1 displays the PM10 loadings and the respective wind speed values throughout 1994. As it can be observed, the values are closely related to the PM10 concentrations; generally decreasing with increasing wind speed. Table 6.1 PM10 and Wind Speed, 1994 1994 January February March April May June July August September October November December PM10 59.7 70.5 65.2 53.9 40.7 61.0 57.5 58.6 49.6 41.1 43.9 63.5 WSP 3.40 3.93 4.25 4.12 4.87 4.22 4.34 3.86 4.38 4.43 4.04 3.40

Although the higher regime lasted until March 1995, the relationship did not stand during this latter period. Apart from the threshold value just mentioned, the size of the particles may affect the ability of the wind to suppress their concentration. In general, particles smaller than 2.5 m are less vulnerable to wind speed as they tend to be associated with combustion sources, while the larger range are composed of crustal elements and dust. The later being more susceptible to suspension by the wind, the relationship between PM2.5, PM10 and wind speed then changes to having a positive relationship between the larger particles and increasing wind speed, and a negative or non existent relationship with the smaller range (Harrison, et al. 1997). Further evidence of the meteorological influence on the particle loadings emanates from the relative humidity and temperature time series, Figures 5.25 and 5.23. When the tri-annual trends of these parameters are analysed together, Figure 5.9, it can be seen that the PM10 concentrations were generally high when the relative humidity and temperature were low and vice versa.

137

This relationship is well expressed during 1994 and 1996, while in 1995 the unusually low relative humidity and PM10 loading during April, disrupted the trend. During this month, the average amount of precipitation was also unusually low, amounting to 8mm instead of 32mm and 37 mm in 1994 and 1996. It was also during this month, that the el nio effect disappeared, and the wind speed returned to its normal values, affecting the overall meteorology of the valley and the concentration of pollutants within the troposphere. No further conclusions can be drawn without weather charts or other indicators of atmospheric stability. However, based on the decrease in wind speed, relative humidity and pluvial precipitation, an argument could be made for a high-pressure system dominating the atmosphere during April of 1995. In an area like the MCMA, with a definite rainy season, the relationship mentioned above becomes important, as the concentration of PM10 will be tied to the seasonal changes of relative humidity, temperature and pluvial precipitation (DDF, 1996). From all the information presented here, it can be stated that the PM10 concentrations were greatly affected by relative humidity, temperature and, under the higher regime, by wind speed.

6.12 O3
The ozone tri-annual time series show a definite seasonal variation, although different to the rest of the pollutants, but consistent with its photochemical origin, shown in Figure 5.10. The concentration was relatively high during the late winter with an average concentration ranging from 0.040 to 0.051 ppm during January and February of 1994, 1995 and 1996. It climaxed in the spring, with a maximum during May and a concentration range in the order of 0.045 to 0.051 ppm. It then descended during the summer and reached a minimum in the autumn-winter period with an average concentration range of 0.031 to 0.048 ppm during December of the three years analysed. As in the case of PM10, this trend was followed in 1995 and 1996 but in 1994, the concentration rose again during the autumn-winter period, as seen in the yearly average plot, Figure 5.11. The increase is partly attributed to the decrease in wind speed from 4.43 to 4.04 and finally to 3.40 m / s from October to December of 1994. The influence of wind speed can be further seen in the O3 monthly averages of the three years analysed, Figures 5.12 to 5.14. Like in the case of the PM10, the concentration distributions of the individual stations in 1994 were much more spread than in the two other years. The influence of meteorology on the tropospheric concentration of ozone can be readily seen in the relative humidity time series, Figure 5.25, and the seasonal distribution of pluvial precipitation, Figure 5.29. The effect emanates from the consumption of ozone and its precursor species.

138

As the relative humidity increased, O3 concentrations generally decreased. Moreover, during the rainy season, the ozone concentrations were always at their lowest, even though it is the time with the most solar irradiation. The consumption of NO2 occurs through its reaction with O3 to produce NO3, reaction 3.56. Further reaction with NO2 leads to the formation of N2O5, a molecule that under normal dry conditions would revert back to NO3 and through photolysis to NO2. Under high relative humidity and conditions leading to rain, however, it may undergo hydrolysis to produce nitric acid via reaction 3.58 and removal by wet or dry deposition. Other removal mechanisms involve the incorporation of HOx radicals into cloud water. In solution, HO2 ionises to form an O2-ion via reaction 3.63, which may then react with O3 in reaction 3.64 to produce OH, OH- and O2. The overall removal process is further enhanced because HO2 in clouds are in direct competition with the gas phase reaction with NO, leading to a further NO NO2 conversion pathway, reaction 3.21. The net loss of O3 through wet depletion (the above processes) is estimated to account for about 20 % of its tropospheric removal processes (PORG, 1997). Considering the time series of both relative humidity and pluvial precipitation and the chemistry just mentioned, it can be stated that the yearly distribution of O3 was strongly affected by the amount of water in the atmosphere (de Icaza, Choularton, 1998). In an area like the MCMA, with a definite rainy season, this becomes important, as the O3 concentration will also be tied to the seasonal distribution of relative humidity and pluvial precipitation. The influence of temperature on the concentration of O3 is not direct, but arises from its association with solar radiation. As it can be seen in both time series, Figures 5.10 and 5.23, the climax of both parameters occurs during May of each year. Further associations are not evident from the time series presented in the previous chapter.

6.13 Meteorology
As it is quite evident from the preceding discussion, the influence of meteorology on the concentration and distribution of PM10 and O3 in the MCMA, and generally in most urban agglomerations, is paramount. The level of influence has already been discussed throughout the last two sections, but to conclude briefly, the following discussion is included. The impact of temperature can be readily seen in PM10 time series and, as explained, indirectly through the O3 trends. As the atmosphere gets warmer, the mixed layer rises to higher altitudes, allowing more ventilation and increasing the turbulence within the boundary layer. Moreover, in a city like the MCMA, that is situated in a valley almost completely surrounded by high mountains, temperature influences the height of the mixed layer by breaking the severe thermal inversions layers experienced throughout the year and especially during the colder months of the year. These inversion layers usually occur during the mornings (most of the time breaking up around midday) at

139

a time when large amounts of the transport related pollutants are emitted, during the peak hours of traffic (de Icaza, Choularton, 1998). Due to the atmospheric stability associated to this condition, pollutants are trapped close to the surface. Unfortunately weather logs were not readily available to illustrate the influence of atmospheric stability and the thermal inversion data was highly sporadic and not very reliable. The effect of relative humidity and pluvial precipitation was to reduce the concentration of pollutants within the MCMA. As it can be seen in their respective time series, both O3 and PM10 had lower loadings during the periods of high relative humidity and abundant rain, that is, during the summer months. As stated, in a city like the MCMA with definite seasonal changes, the pollutant trends affected by precipitation are tightly linked to the seasonally distributed rain. The impact of wind speed on the concentration of PM10 and O3 was clearly marked by the unusual weather generated by el nio phenomenon. During 1994, wind speeds were two to three times the norm and these induced severe changes in the concentration of both pollutants. In the case of the particles, it generally produced lower loadings, especially during the winter period. Ozone, on the other hand, showed increased concentrations during the higher regime but a negative relationship with increasing wind speeds, especially in the last part of the year, the autumnwinter period, shown in Figure 5.11. Overall, the time series analyses show that the distribution of the concentrations of both PM10 and O3 were closely related to the meteorology of the valley. Assuming that the emissions remained relatively constant, it was the stochastic nature of atmospheric processes that drove the changes in their respective concentrations and seasonal distributions.Unfortunately, only the basic meteorological parameters were readily available and thus only simple associations and conclusions were drawn. To enhance the above analysis, pollutant association will be considered and related to the information available from the monitoring networks

6.2 Pollutant Associations


Among the many different ways of classifying pollutants, their nature or origin is important as it helps to predict their concentration within the troposphere. Pollutants emitted directly into the atmosphere are usually termed primary and their concentration depends on their emission sources and their strength. Secondary pollutants, on the other hand, are formed though complex chemical reactions and their concentration depend, among other factors, on the loadings of their precursor species. Atmospheric aerosols can be either primary or secondary, while ozone is only formed through the photolysis of NO2 in the presence of organic species and sunlight. The nature of PM10 in the MCMA and the contribution of secondary aerosols to their mass will be discussed in the following section, while the formation of O3 and the interaction of its precursor species will follow.

140

6.21 PM10
PM10 in urban atmospheres are usually subdivided into one further category, particles having a diameter smaller than 2.5 m or PM2.5. Aerosols in this size spectrum are usually secondary and formed through gas to particle (GTP) processes. They are typically associated with SO2 and NO2 as these two species readily oxidise to form H2SO4 and HNO3 respectively. Their conversion pathways were presented in chapter two; reactions 2.5 to 2.7 for SO2 in the gas phase and 2.10 to 2.12 in the aqueous phase, while reaction 2.8 describes the mechanism for the oxidation of NO2 in the gas phase and 2.13 to 2.16 in the aqueous. The importance of secondary aerosol in an urban airshed depends on the proportion that these represent on the overall PM10. In the MCMA, several studies have been conducted to show that they represent a relatively small part of the overall mass of particles. An official inventory of pollutants undertaken during 1994 (shown in Table 5.3) estimated that 94.2% of all the particulates (by mass) or Total Suspended Particles (TSP) originated from the soil and vegetation, 4.2% were transport related, 0.2% originated form the service industry and 1.2% were attributed to other industry (INE, 1997). These results suggest that most of the particles lay in the coarse mode, above the 2.5 m cut-off point. Further evidence of the size spectra of the particles was found during 1991 by Miranda (1994) in a study of the elemental composition of aerosols within the MCMA. He reported data showing that 18.31% of particles smaller than 15 m lay in the fine fraction, below 2.5 m. Once the proportion has been established, it is important to know the composition of the particles, as these would tend to be associated with the precursor species. In his study, Miranda (1994) attributed 28 3% to soil, 13 + 1% to sulphate, 47 + 2% to organic material and the remaining 10% to other sources. In the coarse mode (2.5m 15 m) however, the profiles showed a 37 2% to soil, 7 + 1% to sulphate, organic content was not measured and the rest to other sources. Although nitrates were not reported, other studies have shown their existence in the aerosol mass of the MCMA. Vega et al (1997) reported that the summation of sulphates and nitrates accounted for 10% of the total PM2.5 mass, while the source profiles of precatalytic converter petrol accounted for 55%, and resuspended dust contributed 34% of the mass. A further study, undertaken by the Metropolitan Commission to Prevent and Control Environmental Pollution in the Valley of Mexico (DDF 1997), showed that the composition of the PM2.5 consisted of 15% sulphates, 16% nitrates, 20% organic carbon and the remaining 49% consisted of other compounds. Although the contribution of PM2.5 to the particle loading was relatively small, they drive the association between PM10, SO2 and NO2. The relationships are depicted in their respective time series, Figures 5.2, 5.15 and 5.17.

141

The time series of SO2 showed no direct relationship to the PM10 tri-annual series. The concentration fluctuated very little and in general at different intervals throughout the three years. NO2 on the other hand, showed more similarities with the PM10 time series. They both had marked seasonal fluctuations and both had similar peaks and troughs, especially during the colder periods of the year. Moreover, they were both related to relative humidity, showing lower concentrations during periods of higher humidity (de Icaza, Choularton, 1998). Figure 5.21 shows the tri-annual trends of the three pollutants and it can be observed that PM10 and NO2 followed closer patterns than PM10 and SO2, suggesting a stronger relationship between the particles and NO2. From these results it can be generally stated that the contribution of secondary aerosol to the mass of PM10 is relatively small and more related to the association between NO2 and PM10 than to the relationship between the particles and SO2.

6.22 O3
Ozone is generated when NO2, in the presence of volatile organic Compounds (VOC), is irradiated with sunlight (reactions 3.10 and 3.9). The relationship between NO2 and O3 is therefore important. On the other hand, ozone is partly removed from the troposphere by titration with NO (reaction 3.12) making the association between O3 and NO also important. The formation and destruction of ozone is discussed in chapter three. The relationship between NO2 and O3 is relatively apparent in the triannual series of both pollutants, Figures 5.17 and 5.10. During periods of high photochemical activity (April to August), ozone concentration follows an inverse relationship with NO2, that is, when the NO2 concentration was low, O3 was high. Under the appropriate conditions, high concentrations of NO2 translate into high concentrations of O3, but the reaction depletes the former species, therefore yielding lower loadings. On the other hand, when irradiation was relatively low, the relationship reversed and high concentrations of NO2 produce large quantities of O3. Figure 5.22 shows the tri-annul time series of O3, NO2 and NO together. This trend was apparent during 1994 and 1996, while in 1995 high loadings of O3 were equalled by low NO2 concentrations throughout the year. From January to August of this year, however, the NO2 concentrations were lower than in 1994 and 1996. This tendency is probably related to a decrease in NOx concentrations as NO showed a similar trend, lower loadings during 1995, as seen in tri-annual trends of the three pollutants (de Icaza, Choularton, 1998). The disruption of the trend was related to the consumption of NO2 to produce O3. As can be seen in Figure 5.11, the concentration of ozone was higher in 1995 than in half of the months of 1994 and higher than most of 1996.

142

As the NOx emissions were lower, smaller quantities of NO2 were photolysed to produce larger quantities of O3, all relative to the concentrations experienced in the two other years. Further evidence of this hypothesis can be observed in the NO yearly trend, Figure 5.20, as it shows that it was also lower, and at periods of high O3 concentrations NO was low, consistent with the chemistry involved (reaction 3.12). The trends of NO and O3 during 1995 display a typical scenario; with high NO concentrations titrating large quantities of O3. In this manner when NO was high, O3 was low and vice versa (as shown in Figure 5.22). The trend just mentioned was also followed during 1994 and1996, but not as pronounced in periods of high photochemistry. Most NOx in urban atmospheres are emitted in the form of NO, as these are formed during combustion of any high temperature process involving air. Although NO is the primary pollutant, it oxidises very efficiently to form NO2. The main source of these processes is the automobile pool of the city. In the MCMA, there are more than 2.7 million vehicles, all of which contribute to the concentration of NO and by association to concentration of NO2 and to O3.In the official inventory of pollutants (DDF 1997), NOx sources were mainly attributed to the transport sector. Over 71 % of all NOx are emitted by motor vehicles, the specific breake down shows private cars contributing 25%, microbuses 12%, Taxis 7% and other transport including buses and lorry trucks accounting for 27%. The remaining processes contributing to NOx emissions include thermoelectric plants with 14%, general industry 11% and services 4%. The concentration of NOx is relatively important as it drives the formation and destruction of O3. The ratio of VOCs to NOx determines the potential strength of ozone formation. If the ratio is high (> 10), a condition known as NOx limited, then the formation is dependent on the availability of NOx. If on the other hand the ratio is low (< 10), then the atmosphere is termed VOC limited and the formation strength is dependent on the availability of VOCs. This relationship can be visualised in the O3 isopleth, Figure 3.1 in chapter three. In this chart, VOC limited conditions tend to fall towards the right of the diagram, while the NOx limited region is on the left. In the MCMA, the ratios are very high (in the range of 13 to 50), and therefore large amounts of NOx are required to satisfy the relative abundance of VOCs (DDF, 1997). Under these conditions, a reduction in NOx would, in theory, produce a reduction in O3 concentrations. Drawing on this information, it can be deduced that during 1995, the relative abundance of VOCs was low, as a reduction in NOx produced larger loadings of ozone.

6.3 Diurnal Analyses


The formation and transformation of pollutants within the troposphere differs throughout the diurnal cycle. Primary pollutants are emitted during the day and night and are dependent, among other factors, on the level of social activity of the region.

143

Photolytic pollutants on the other hand, can only be formed during the day and their concentration is bound to the levels of insolation and in the case of O3 and secondary particles, on the concentration of their precursor species. Diurnal plots were prepared to show the different behaviour of PM10, O3, and their precursors within the atmosphere of the MCMA. These graphs show the day time period lasting from 07 hrs. to 20 hrs., while the night time cycle extends form 21 hrs. to 06 hrs. In the following section, the diurnal behaviour of PM10, O3 and their precursors will be discussed. The particles will be related to SO2 and NO2, while O3 will be discussed in terms of its association with NO2 and NO. The diurnal distribution of the PM10 concentration was fairly normal, showing higher concentrations during the day, Figure 5.32. The distributions of both day and night follow the same seasonal variations, high in the early winter, descending during the spring, reaching a minimum during the summer and then ascending during the autumn and reaching a maximum during the late winter. The difference in concentration between the day and the night was relatively large, with a minimum variation during May of 1996 (2 g/m3) and a maximum value in March of 1996 (36 g/m3). Throughout the year, the diurnal range was relatively large (average 21 g/m3), suggesting that emissions and / or formation occurred preferentially during the day-time, at a time when social activity is at its highest. The diurnal trends of NO2 and SO2, Figures 5.35 and 5.33, also showed a typical distribution, with higher concentrations in the day. While both pollutants follow their seasonal variability throughout the day time and night time periods, SO2 concentrations had a higher range during the summer of 1995 with much lower values during the night. The variation exhibited by SO2 ranged from 0.001 ppmv in August 1996 to 0.009 ppmv during June 1995. It must be noted that the negative value was the only one in the time series and that the average difference in concentration was 0.005 ppmv. NO2 on the other hand, showed a diurnal variation ranging from 0.000 ppmv in February 1994 to 0.017 ppmv during January 1996. As in the case of SO2, the zero value was exceptional and the average difference was 0.009 ppmv. The diurnal plots of these pollutants did not show any association between them and the particles. They have minimum and maximum ranges at different times throughout the three years and their overall variation was dissimilar. This is explained by the relatively small proportion of particles falling in the secondary category. The diurnal variation exhibited by ozone was very pronounced, with an average difference of 0.042 ppmv, shown in Figure 5.34. Values during the day were much higher than in the night, consistent with its photochemical origin, as it is only produced in the presence of sunlight. The variation ranged from 0.032 ppmv in June 1995 to 0.058 ppmv during March of the same year.

144

Notwithstanding the difference in concentration between the day and the night, the seasonal distribution of both periods was the same throughout the year. The loadings began high in the early winter, reached a maximum during the spring, descended in the summer and increased again in the autumn and late winter. The climax during May was evident on both the night and day time trends. The relatively large diurnal variation exhibited by ozone can be explained through its night time chemistry. O3 reacts with NO2 to produce NO3 through reaction 3.56, which then reacts with another molecule of NO 2 to produce N 2 O 5 via reaction 3.57. These reactions occur mainly after dusk, as NO3 is rapidly photolysed back to NO2 in the presence of sunlight, reaction 3.60. This mechanism, as discussed before, is important in the atmosphere of the MCMA and it provides an important sink for O3, especially during the summer months when pluvial precipitation is abundant. Further ozone sinks include its reaction with NO, to replenish the NO2 concentration via reaction 3.12. All of these relationships are well represented in the respective diurnal plots, Figures 5.34 to 5.36. NO2 showed some variation but not as much as O3, while the NO concentrations were very similar during the day and the night time periods. This is because NO requires the presence of O3 (which is low during the night) to replenish the NO2 concentration, while NO2 reacts with O3 to produce the nitrate radical throughout the night. In this manner, NO2 and O3 are reacting during the night, and NO only reacts during the day. From the above information it can be concluded that the diurnal variation of O3, NO and NO2 was related to the level of insolation. Although NO was mostly emitted during the day, its removal from the atmosphere was just as efficient during the day and the night. NO2 on the other hand, reacted with O3 throughout the night and therefore its concentration fluctuated during the diurnal cycle. O3 was removed from the atmosphere by night and its production only occurred during the daylight hours, therefore showing the largest range in diurnal difference.

6.4 Statistical Analyses


The conditions in which pollutants form are varied and can be very complex. In this section the relationship between PM10, O3, their precursors and the relevant meteorological parameters will be quantified by applying Ordinary Least Square (OLS) and multiple regression analysis. The actual statistical analysis cannot be tested for significance because the Durbin Watson test shows that the error terms of the regressions, also referred as the residuals, are autocorrelated.

145

In other words, the residuals are not statistically independent, an assumption of the OLS regression. The effect of this violation is that the confidence intervals around the estimators are very large, thus they are inefficient and the t and F tests are invalid. The estimates, however, are unbiased. The statistic numbers reported in Table 5.4 reveal that, according to this test, all the residuals of the regressions calculated are autocorrelated as the estimates lay below the lower boundary stipulated by the test, 1.46. None the less, the estimates are unbiased and the relationship between the parameters is still representative. Ordinary Least Square (OLS) regressions were performed to the data corresponding to the last year studied, 1996. The analyses corroborated some of the previous hypotheses and in a quantitative manner suggested possible associations between PM10, O3, their precursors and meteorological parameters. The results of the regression analyses are detailed in Table 5.4. As reported earlier, 18.31% of the PM15 lay below the 2.5 m size spectra (Miranda, 1994), suggesting that these are secondary and that their formation may be associated with SO2 and NO2. The Coefficient of Determination (COD or R2) of the relationship between SO2 and PM10 showed that they were not well correlated, R2 = 0.059, while PM 10 and NO 2 had a relatively strong relationship, R 2 = 0.31. Temperature was not very important, R2 = 0.002, but relative humidity was well correlated, with an R2 equal to 0.31. From these results, multiple regression was applied to the two factors showing a strong association, relative humidity and NO2. Taken together, they showed a relatively high COD, R2 = 0.44, suggesting that the association between these two parameters influenced the concentration of PM10 in the troposphere of the MCMA. Furthermore, by inspecting the statistical plots, Figures 5.40 and 5.38, it can be seen that relative humidity is negatively correlated with PM10, and that the relationship between NO2 and PM10 is positive. These facts suggest that NO2 was involved in the formation while relative humidity in the removal of PM10. The regression analyses undertaken to quantify the associations with O3 reinforce several hypotheses proposed earlier. Its relationship with relative humidity was strong, R2 = 0.20, while temperature showed a weak correlation, R2 = 0.084. These results corroborate the conclusion posed before, that temperature was not directly affecting the concentration of O3, but relative humidity was involved in its removal from the troposphere. This is further seen in the statistical plots where it can be observed that the relationship between O3 and relative humidity was negative, with a slope of 0.003, Figure 5.44. Further conclusions can be drawn from the associations between O3 and the species involved in its formation and titration.

146

Ozone was well correlated with NO2, R2 = 0.23, while the relationship with NO was poor, R2 = 0.0002. From this difference in correlation, it can be stated that the concentration of O3 in the troposphere of the MCMA was dictated by the photolysis of NO2 and not by its titration with NO. Evidence of these relationships can also be seen in their respective diurnal plots, where the concentration of NO did not fluctuate between day and night and where NO2 and O3 showed relatively large fluctuations, Figures 5.34, 5.35 and 5.36. As in the case of the PM10, the parameters showing a strong relationship with O3 were analysed together via multiple regression. In combination, relative humidity and NO2 had a relatively high COD, R2 = 0.49, suggesting that these two parameters were partly controlling the concentration of O3 in the troposphere of the MCMA. Although all these values are relatively subjective, as they cannot be tested for significance, they do quantify the relationships. They show, among other factors, that secondary PM10 (below the 2.5 m cut off point) were mainly produced by the oxidation of NO2, while SO2 played a smaller role. The results also suggest that relative humidity influenced the loading of both PM10 and O3 in the troposphere of the MCMA. Furthermore, the concentration of O3 was controlled by the photolytic reaction of NO2 and not by its titration with NO. The formation conditions of PM10 and O3 can be further assessed by inspecting the statistical plots showing their relationship with their respective precursors and meteorological parameters. Figures 5.37 to 5.40 display the association between PM10 and SO2, NO2, temperature and relative humidity, while the relationships of O3 with NO2, NO and the same meteorological parameters are depicted in Figures 5.41 to 5.44. According to the statistical plots, PM10 were present when the temperature was around 15C to 20C, the relative humidity ranged from 30% to 80% and the concentration of SO2 and NO2 ranged from 0.010ppmv0.020ppmv and 0.015ppmv0.080ppmv, respectively. The highest concentration of PM10 occurred when the temperature was 17C, relative humidity was 65%, and the concentration of SO2 and NO2 were 0.010 ppmv and 0.048 ppmv, respectively. The plots displaying the O3 relationships show that it was present when the temperature was between 10C 22C, the relative humidity ranged from 20% 80% and the concentration of NO2 and NO were around 0.020 0.080 and 0.015 0.114 ppmv respectively. The highest concentration of O3 occurred when the temperature was 20 C, the relative humidity around 27%, the NO2 concentration was 0.053 ppmv and finally the NO loading was 0.051 ppmv.

147

All these values closely agree with the theoretical formation mechanisms as the temperature was relatively high, meaning that sunlight was available, and the NO2 contribution was also present.

6.5 Wind Analyses


In order to finish the analysis of the formation and transformation of PM10 and O3 in the atmosphere of the MCMA, prevailing wind and pollutant roses were calculated. The purpose is to attempt an assessment of the interactions between the pollutants and to gain an understanding of their possible sources. As before, PM10 will be related to SO2 and NO2, while O3 will be assessed in terms of its association with NO2 and NO. The prevailing wind will be taken into consideration mainly because the MCMA has a relatively large industrial zone in the northern part of the area. According to the frequency of occurrence, the prevailing wind direction was from the north in all the three years analysed and in the agglomerated frequencies, Figures 5.45 and 5.46 respectively. As stated before, a large proportion of the industrial base of the MCMA is located in the north. Furthermore, a large thermoelectric plant is also situated in the north and therefore the pollutants emitted from these are being transported throughout the city. These can be readily seen in the respective pollutant roses, Figures 5.47, 5.48 and 5.51, where the highest loadings of PM10, SO2 and NO originated from the northern part of the area, in the TLA and XAL stations. The graphs showing concentration frequencies for PM10 illustrate their relationship with NO 2 and SO 2 , Figures 5.47, 5.49 and 5.48. The roses of the three pollutants show that the concentrations were spread, indicating that the sources were varied and, in general, not very localised. The loadings of NO2 and PM10 were in close agreement, with similar sources in most of the stations, while SO2 showed little indication of being related. In their respective figures, it can be seen that PM10 and NO2 had a strong relationship in most of the stations. In TLA, the high concentrations (in descending order) came from the southeast, east and south. In XAL, the relationship was the same, coming from the south-east and east, while in MER the concentrations were high throughout. The stations in the south, however, showed weaker relationships with the highest concentrations coming from the south in PED and there was no indication of any association in CES. The direction of the highest concentrations from PM10 and NO2 further corroborate the strength of their relationship, while SO2, as in the rest of the analysis, had a secondary role. Table 6.2 displays the concentration in three of the stations in the MCMA.

148

They follow the trajectory from north to south, with XAL being in the northeast, MER in the city centre and PED in the south-west. The trajectory of the PM10, SO2 and NO, as can be seen in the corresponding values, is from north to south, consistent with the prevailing wind direction and with large industrial and power generation sources in the northern part of the city. Table 6.2 Pollutant Trajectory during 1994-96 in three stations Year 1994 1995 1996 Average 1994 1995 1996 Average 1994 1995 1996 Average 1994 1995 1996 Average 1994 1995 1996 Average XAL 43.91 100.94 122.37 66.81 0.024 0.017 0.020 0.020 0.037 0.030 0.039 0.035 0.043 0.039 0.028 0.037 0.053 0.056 0.070 0.060 MER 49.08 40.73 59.95 37.44 0.019 0.015 0.020 0.018 0.052 0.037 0.043 0.044 0.046 0.044 0.033 0.041 0.044 0.03 0.051 0.045 PED 45.75 39.61 48.94 33.58 0.015 0.014 0.012 0.014 0.042 0.032 0.037 0.037 0.050 0.049 0.045 0.048 0.022 0.022 0.024 0.023

PM10 (g/m3)

SO2 (ppmv)

NO2 (ppmv)

O3 (ppmv)

NO (ppmv)

The relationship between O3 and NO2 and NO can also be seen in their respective rose charts, Figures 5.50, 5.49 and 5.51. The highest loading of O3 occurred at the PED station, in the south-west of the city, and came from the south and the south-east, while NO2 also showed high concentrations in the same direction. In this station, the relationship with NO was not evident. The rest of the stations showed some type of relationship, either with NO2 or with NO. At the station in the north-west, TLA, the association between ozone and NO2 was strong from the south, while NO showed its highest loading from the west and north-west, the same direction where O3 was relatively low.

149

In the north-east station, XAL, there was no evidence of a relationship with NO2, but NO was again high where O3 was low, from the north and northeast. At the city centre, in the MER station, the loadings were high throughout, although O3 showed a higher loading from the north-east, the direction with the lowest NO concentration. The relationship of NO2 with ozone was again strong in the south-east of the city, in the CES station. The loading of NO2 and O3 were high when coming from the north-east, north and north-west, while NO was relatively low in the north-west. As can be seen in their respective figures, the concentrations of O3 and NO2 coincided in most of the stations, displaying the role of the photolysis of NO2 to form O3. The relationship between O3 and NO was also evident, high loadings of NO titrated large amounts of O3, hence the directions with relatively low NO propitiated high ozone concentrations. In terms of the values reported in Table 6.2, the concentration of O3 increased from south to north, being highest in the PED station, lower in MER and lowest at XAL. The NO2 concentrations showed that the strongest sources were in the city centre at the MER station, while NO followed the direction mentioned before, north to south. These values are consistent with the chemistry involved and the wind patterns of the city. O3 was probably produced in the city centre, where the highest loadings of NO2 were registered, and was transported to the south in the urban plume. As seen in Table 6.2, the NO2 concentrations were highest in the MER station and then decrease towards the south-west, in the PED station. From all the information presented above, several conclusions can be corroborated and others stated. PM10, NO and SO2 had strong emission sources in the northern part of the city, where a large proportion of the industrial base and a power station are located. The loading of these pollutants increased in north to south trajectory, consistent with the prevailing wind direction. PM10 were relatively well associated with NO2 as their highest loadings came from the same direction, while SO2 showed little or no association. Ozone was mainly produced in the city centre and transported to the southwest of the city, where the highest concentrations were registered. The association between O3 and NO2 was evident in most of the stations and the relationship with NO was also present in most stations.

6.6 Conclusion
Throughout this chapter, the information presented in the results of the MCMA was analysed and discussed. Time series analyses were carried out to understand the influence of meteorology on the concentration of PM 10 and O 3 . Pollutant associations were examined and then quantified through simple statistical tests. The influence of social activity and the level of insolation were also assessed through plots showing the diurnal distribution of the particles, ozone and their precursor species.

150

Finally, wind direction was incorporated into the overall analysis by drawing conclusions on the relationships established in the previous sections. The time series of PM10 revealed that their concentration was greatly affected by temperature, relative humidity and the distribution and amount of pluvial precipitation. The loadings of the particles were highly seasonal during the three years analysed and, in general, they agreed with patterns of relative humidity, temperature and rain. The concentrations were high during the winter when temperature, relative humidity and rain were at their lowest and low during the summer when the three meteorological parameters were at their highest. The influence of wind speed was large under the higher regime, exhibited during 1994. In this year, the oceanic effect of el nio was present and it affected the airshed of the MCMA by producing wind speeds two to three times higher than the norm. During this year, the loadings of the particles were lower, and their sources more dispersed than in the two latter years. The influence of meteorology was also evident in the O3 time series. They showed that the concentration was also seasonally distributed, but related to its photochemical origin. O3 loadings climaxed in May of the three years were high during the early winter (January and February), reached a peak in the spring, descended in the summer and reached a minimum the autumn-winter period. This trend was consistent with the period of highest photochemical activity as the temperature reached its maximum at the same time. Moreover, O3 concentrations decreased at the same time that relative humidity and pluvial precipitation increased, during the summer. The influence of wind speed was also marked during autumn-winter period of 1994, with higher ozone concentrations being related to decreasing wind speeds. In the pollutant associations, less than 20% of the particles by mass were determined to fall below the 2.5 m size range, suggesting that only a small proportion of the PM10 were secondary and related to the oxidation of SO2 and NO2. Through their intra-annual distributions, in the time series of the three pollutants, it was concluded that the patterns followed by NO2 were much closer to the particles than the distributions of SO2, suggesting a closer association between NO2 and PM10. The O3 associations showed that its concentration is inversely proportional to the concentration of NO2 during periods of high photochemistry and proportional during the rest of the year. NO had a typical association with O3; high NO titrated large quantities of O3. The diurnal analyses showed that PM10 were preferentially emitted and / or produced during the day time period, at the time when social activity is at its highest. The diurnal distributions of neither NO2 nor SO2 were well related to the particles, possibly indicating the small proportion of secondary particles within the PM10. Ozone, as expected, showed a large diurnal range, which was consistent with its photochemical origin.

151

NO2, as well, showed a marked variation between day and night while NO distributions displayed little variation between the two periods. The diurnal ranges were explained according to their inherent chemistry, with NO2 and O 3 reacting during the night and NO being kept constant. The statistical analyses helped to quantify the associations found earlier. They showed a strong positive relationship between NO2 and PM10, while a strong and negative relationship between relative humidity and the particles. These results suggested that secondary particles were produced preferentially by the oxidation NO2 and removed by rain under high relative humidity. The rest of the parameters were not well correlated to the particles. O3 was also negatively correlated with relative humidity and positively associated with NO2, suggesting that it was removed by the rain associated with the high relative humidity and produced through the photolysis of NO2. The correlation of O3 and NO indicated that the concentration of ozone in the atmosphere of the MCMA was dictated by the photolytic reaction of NO2 and not by its titration with NO. Finally, wind direction analyses showed that large emissions of PM10, NO and SO2 originated form the north, where a large proportion of the industrial base and a power generation plant are located in the MCMA. The loadings of these pollutants dispersed throughout the city in the trajectory of the urban plume, showing diminishing concentrations in a southerly direction. Furthermore, the associations between NO2 and the particles and between O3 and NO2 were corroborated as the highest concentrations originated from the same direction. The relationship between O3 and NO was also evident, with large quantities of ozone originating in the direction of lower NO loadings. From this analyses, it was also shown that most of the ozone was produced in the city centre and transported to the south. The concentration of O3 was highest in the south-west and decreased in a northerly direction, opposite to the NO loadings. Overall, the results show that PM10 in the MCMA are mostly primary in nature and their concentrations were affected by temperature, relative humidity, rain and under the higher regime by wind speed. The secondary fraction was mainly attributed to the oxidation of NO2 and little or no association was found between SO2 and the particles. PM10 were mostly emitted during the day and in the northern part of the city. Ozone on the other hand was produced in the city centre and its concentration was also affected by relative humidity, rain and under the higher regime by wind speed. The concentration of ozone was controlled by the photolysis of NO2, although NO reduced its concentration during the day.

152

Results from the GMA


Chapter Seven

7.1 Introduction
As in the case of the MCMA, time series were calculated in monthly-hourly periods during 1997. Only one year of data was analysed for the GMA as the network is relatively new, and longer periods were not available for all the pollutants. Five of the seven stations within the area monitor all the pollutants consistently, four measure PM10 and precursors and four report ozone and its associated pollutants (one station in common). The stations and the pollutants that they measure are represented in Table 7.1 and are displayed in the following map, Figure 7.1. Table 7.1 Monitoring Stations and Parameters Monitored Station Bolton Bury Piccadilly Stockport South Pollutant PM 1 0 , SO 2 , O 3 , NO, NO 2 , NOx PM 1 0 , SO 2 , O 3 , NO, NO 2 , NOx PM 1 0 , SO 2 , O 3 , NO, NO 2 , NOx PM10, SO2, NO, NO2, NOx O3, NO, NO2, NOx and Met. Parameters.

Figure 7.1 Map of the GMA with the Monitoring Stations (Adapted from Barlow, 1995)

GMA City Boundary Monitoring Station

Rochdale Bolton
Bolton

Bury
Bury Roadside

Oldham

Wigan

Salford
Salford Eccles

Trafford

Manchester Picadilly Tameside


Town Hall

Stockport
South
Stockport

155

Annual trends were calculated for PM10, SO2, O3, NO and NO2 in order to understand their associations and distribution throughout the year. The basic meteorological parameters were also graphed and will be used to illustrate their influence on the formation and transformation of PM 10 and O 3 . Diurnal charts showing the distribution of the same pollutants are included to show the different levels of activity within the area and the influence of insolation on the tropospheric concentration of secondary pollutants. The aim of this chapter is to present the main results of the pollutant trends, their diurnal behaviour, associations and the probable sources of PM10 and O3. All the work presented here will be placed into context and analysed in the subsequent discussion of these results.

7.2 Time Series


As mentioned, annual trends were calculated in monthly-hourly periods during 1997. These show the main distribution of the concentration of the pollutants throughout the year.

7.21 PM10 Time Series


The yearly average concentration of PM10 is presented in Figure 7.2. This graph shows a relatively high concentration during the month of January, followed by substantially lower concentration during February. The rest of the year shows a small trend, with decreasing loadings during the spring, lower in the summer and slightly higher in the autumn and winter periods. The concentrations in each of the stations analysed are presented in Figure 7.3. Figure 7.2 PM10 Yearly Average, 1997
Ave. (24 ug/m3) Ave. (22 ug/m3, excluding Bury)

40 35 30 25

ug/m3

20 15 10 5 0 Jan Feb Mar Apr May Jun Jul Ago Sep Oct Nov Dec

156

Figure 7.3 PM10 Monthly Average, 1997


Piccadilly Bolton Bury Salford Stockport

60

50

40

ug/m3

30

20

10

0 Jan Feb Mar Apr May Jun Jul Ago Sep Oct Nov Dec

In this plot, it can be seen that the PM10 concentration at Bury, the station to the north-west of the region, is constantly reporting higher values than the rest.

7.22 SO2 Time Series


Sulphur dioxide is a major component of gas to particle conversion and as such, it influences the concentration of PM10. The plot in Figure 7.4 shows the SO2 time series during 1997 and it can be seen that it is very similar to the PM10 distribution, high in January and low in February. In the rest of the months, there is a slight trend; lower during the summer and higher in the winter. As in the case with the PM10, the values reported by the Bury are considerably higher than in the rest of the stations. As will be explained latter in the discussion chapter, this station is next to a motorway with a daily usage of more than 200,000 vehicles per day.

157

Figure 7.4 SO2 Yearly Average, 1997

Ave. (8 ppb) Ave. ( excluding Bury, 6 ppb)

16 14 12 10

ppb

8 6 4 2 0 Jan Feb Mar Apr May Jun Jul Ago Sep Oct Nov Dec

The relationship between PM10, NO2 and SO2 is very important as secondary particles depend on the concentration of the two latter ones. Figure 7.5 displays this relationship.

Figure 7.5 PM10, NO2 and SO2 Yearly Averages, 1997

35 30 25 20

35 30 25 20

ug/m3

ppb

15 10 5 0 Jan Feb Mar Apr May Jun Jul Ago Sep Oct Nov Dec

15 10 5 0

NO2 Ave. 97 (Excluding Bury, 21 ppb PM10 Ave. 22ug/m3, excluding Bury

SO2 Ave. (excluding Bury, 6 ppb)

158

7.23 Ozone Time Series


Ozone is a secondary pollutant produced by the photolysis of NO2 and titrated by reaction with NO. The trend followed by O3 is shown in Figure 7.6, where it can be seen that the concentration follows a definite seasonal trend; high in the spring and summer, while lower during the autumn and winter. The values reported at Bury are much lower than in the rest of the station. This is related to the high NO loadings reported at the same station. Figure 7.6 O3 Yearly Average, 1997
O3 Ave. (13 ppb) O3 Ave. ( 14 ppb excluding Bury)

25

20

15

ppb
10 5 0 Jan Feb Mar Apr May Jun Jul Ago Sep Oct Nov Dec

The lower and higher values reported during January and February are most likely related to synoptic weather conditions experienced during these two months of the year. The monthly average concentrations in each station are shown in Figure 7.7.

159

Figure 7.7 O3 Monthly Average, 1997

30

Bolton Bury Piccadilly Salford South

25

20

ppb

15

10

0 Jan Feb Mar Apr May Jun Jul Ago Sep Oct Nov Dec

The plot above illustrates the influence that NO has on the tropospheric concentration of O3. As NO increases, ozone is titrated to lower tropospheric concentrations. The ozone averages at Bury are comparatively lower than the rest of the stations, as can be expected. The rest of the stations show similar concentration distributions. As the stations are located nearer to the city centre, or in locations with large emissions of nitric oxide, the concentration of O3 will be lower. This can be readily seen in the figure above where Bury, next to a motorway, has lower concentrations than Bolton or South, both suburban locations.

7.24 NOx (NO and NO2) Time Series


The trend for NO is shown in Figure 7.8, where it can be observed that it follows the O 3 time series very closely, but in the opposite direction. It is high during January and low in February, as the rest of the primary pollutants. The concentration at Bury reported some of the largest NO loadings, as its location is affected by strong sources of motor exhaust pollutants.

160

Figure 7.8 NO Yearly Average, 1997


Ave 97 (41 ppb) Ave 97 (excluding Bury, 26 ppb)

120

100

80

ppb

60

40

20

0 Jan Feb Mar Apr May Jun Jul Ago Sep Oct Nov Dec

The trends for NO2 reflect the relationship between NO and O3, as NO2 is required to produce O3, while NO titrates O3 to produce NO2 once more (Equations 3.9 to 3.12 in the Ozone Formation chapter). The concentration starts high in January and lowers during February following the NO trend but avoiding the extreme values observed in that distribution. Figure 7.9 portrays the NO2 yearly average concentration, with a secondary curve illustrating the trend without the Bury concentrations. Figure 7.9 NO2 Yearly Average, 1997
Ave 97 (24 ppb) Ave 97 (excluding Bury, 21 ppb)

35 30 25 20

ppb
15 10 5 0 Jan Feb Mar Apr May Jun Jul Ago Sep Oct Nov Dec

161

The interrelationship between NO, NO 2 and O 3 can be observed in Figure 7.10. The average concentration of NO is the highest of the three, illustrating the ozone destruction nature of the atmosphere throughout the GMA. When the concentration of the nitrogen oxides (NO and NO2) is high, the corresponding ozone value is low and visa versa. This is true in all the months except in June when the value for NO is lower and the NO2 average is the same as O3. This low value is, however, only related to the NO concentration reported at Bury. As can be seen in Figure 7.11, the relationship between the three pollutants maintains a more normal trend when Bury is excluded from the trends.

Figure 7.10 NOx and Ozone Trends, 1997


NO Ave 97 (41 ppb) NO2 Ave. 97 (24 ppb) O3 Ave. 97 (13 ppb)

120

100

80

ppb

60

40

20

0 Jan Feb Mar Apr May Jun Jul Ago Sep Oct Nov Dec

162

Figure 7.11 NOx and Ozone Trends (excluding Bury), 1997


O3 Ave (14 ppb excluding Bury) NO2 Ave. 97 (24 ppb) O3 Ave. 97 (13 ppb)

70 60 50 40

ppb

30 20 10 0 Jan Feb Mar Apr May Jun Jul Ago Sep Oct Nov Dec

7.25 Meteorological Parameters


Pollutants concentrations within the atmosphere of an area are highly dependent on the prevailing meteorological parameters. In this section, the time series of temperature, relative humidity and wind speed will be presented. The temperature profile shows a relatively normal distribution, high in the summer, lower in the equinoxes and lowest during the winter. Figure 7.12 displays the temperature trend in the GMA during 1997. Figure 7.12 Temperature Yearly Average, 1997
Temp. (Ave. 10.5 Deg.C )

20 18 16 14 12

Deg. C.

10 8 6 4 2 0 Jan Feb Mar Apr May Jun Jul Ago Sep Oct Nov Dec

163

Relative humidity affects among other factors the formation and residence time of several pollutants in the atmosphere, especially PM10. Figure 7.13 shows the average distribution of relative humidity throughout the year. Figure 7.13 Relative Humidity Yearly Average, 1997
RH (Ave. 79.1 %)

100 90 80 70 60

ppb

50 40 30 20 10 0 Jan Feb Mar Apr May Jun Jul Ago Sep Oct Nov Dec

The concentration of pollutants within the troposphere is highly dependent on the wind speed of the area. During 1997, the average wind speed in the GMA fluctuated from 6 to 13 metres per second, accounting for different periods of high and low pressure systems. Figure 7.14 shows the time series throughout the year. Figure 7.14 Wind Speed Yearly Average, 1997
Wind Speed (Ave. 8 m/s)

14 12 10 8

m/s

6 4 2 0 Jan Feb Mar Apr May Jun Jul Ago Sep Oct Nov Dec

164

7.26 Rainfall Time Series


In order to finish the time series analyses, precipitation data must be analysed. During 1997, pluvial precipitation amounted to 765.1 mm and it was distributed in 186 days. Also important is the distribution of the rain in the diurnal cycle: a total of 375.9 mm fell during the night while 389.2 mm occurred in the daytime. Figures 7.15 and 7.16 show the daily amount of rainfall in 1997 and the distribution of the monthly average, respectively. Figure 7.15 Daily Rainfall, 1997
30

25

20

mm

15

10

0
1 5 1 /9 /0 7 2 9 1 /9 /0 7 1 2 1 /9 /0 7 2 6 2 /9 /0 7 1 2 2 /9 /0 7 2 6 3 /9 /0 7 0 9 3 /9 /0 7 2 3 4 /9 /0 7 0 7 4 /9 /0 7 21 5/ /0 9 7 0 4 5 /9 /0 7 1 8 6 /9 /0 7 0 2 6 /9 /0 7 1 6 7 /9 /0 7 3 0 7 /9 /0 7 1 8 7 /9 /0 7 27 8/ /0 9 7 1 0 8 /9 /0 7 2 4 9 /9 /0 7 0 8 9 /9 /1 7 2 2 0 /9 /1 7 05 0/ /1 9 7 19 1/ /1 9 7 0 3 1 /9 /1 7 1 7 2 /9 /1 7 3 1 2 /9 /1 7 2/ 97 /0 01

Figure 7.16 Monthly Rainfall, 1997


Day Rainfall Night Rainfall Monthly Rainfall

120

100

80

mm

60

40

20

0 Jan Feb Mar Apr May Jun Jul Ago Sep Oct Nov Dec

165

7.3 Diurnal Analyses


Diurnal variation charts were prepared for the purpose of determining the influence of social activity, temperature and insolation on the ambient concentration of O3 and PM10. As in the case of the MCMA, a large proportion of the atmospheric pollutants is derived from the transport sector of the economy. This translates into having large emissions of sulphur dioxide, nitric oxide, nitrogen dioxide, aerosols and through chemical and photochemical processes, high concentrations of ozone. Most of these emissions will also occur during the daytime hours of the day, therefore increasing (or decreasing) the concentration of pollutants during the peak hours of social activity. Taking into consideration the relatively northerly location of the area, the resulting large variation of daylight hours and the large differences in temperature between winter and summer, two different sets of analyses were performed. During the winter, in January and February, when the amount of daylight hours are relatively scarce, the day-time hours were selected as 10:00 to 15:00 and the night-time as 18:00 to 06:00 hrs. In the summer, however, the hours were changed to reflect the large amount of daylight to the following schedule; 08:00 to 20:00 hrs for the day and 22:00 to 03:00 during the night. Also important in these analyses is the fact that some of the monitoring stations do not measure all the pollutants during the two regimens just mentioned. To account for this problem, only the stations that consistently report all the necessary concentrations are included in the diurnal averages, and are mentioned at the left-hand corner of each graph. As in the past analyses, the main pollutants of interest are represented in the following plots. Starting with PM10, the graphs show a relatively low variation between the day and night, but a relatively higher difference between summer and winter. The plots showing the variations of PM10 concentrations for winter and summer are presented in Figures 7.17 and 7.18 respectively.

166

Figure 7.17 PM10 Diurnal Average, Winter 1997


Day time 10-15 hrs. Night time 18-06 hrs.

120

100

80

ug/m3

60

40

20

0
01/01/97 08/01/97 15/01/97 22/01/97 29/01/97 05/02/97 12/02/97 19/02/97 26/02/97

Figure 7.18 PM10 Diurnal Average, Summer 1997


Piccadilly Bolton Bury Roadside Stockport Day time 08-20 hrs. Night time 22-03 hrs.

80 70 60 50

ug/m3

40 30 20 10 0
97 97 97 97 97 97 97 97 97 97 97 97 97 8/ /0 31 6/ 6/ 6/ 6/ 6/ 7/ 7/ 7/ 7/ 8/ 8/ 8/ /0 /0 /0 /0 /0 /0 /0 /0 /0 /0 /0 /0 /0 8/ 97

The origin of particulates in the GMA is most likely secondary. This can be seen in the annual trends, but in the diurnal averages it becomes noticeable as well. The diurnal plots for SO2 are shown in Figures 7.19 and 7.20, during winter and summer, respectively.

01

08

15

22

29

06

13

167

20

27

03

10

17

24

Figure 7.19 SO2 Diurnal Average, Winter 1997


Piccadilly Bolton Bury Roadside Stockport

Day time 10-15 hrs. Night time 18-06 hrs.

40 35 30 25

ppb

20 15 10 5 0
01/01/97 08/01/97 15/01/97 22/01/97 29/01/97 05/02/97 12/02/97 19/02/97 26/02/97

Figure 7.20 SO2 Diurnal Average, Summer 1997


Piccadilly Bolton Bury Roadside Stockport

Day time 08-20 hrs. Night time 22-03 hrs.

20 18 16 14 12

ug/m3

10 8 6 4 2 0
97 97 97 97 97 97 97 97 97 97 97 97 97 8/ /0 31 6/ 6/ 6/ 6/ 6/ 7/ 7/ 7/ 7/ 8/ 8/ 8/ /0 /0 /0 /0 /0 /0 /0 /0 /0 /0 /0 /0 /0 8/ 97

The variation between the ozone concentration during the summer and winter, and between day and night is presented in the next two graphs, Figures 7.21 and 7.22 for winter and summer respectively.

01

08

15

22

29

06

13

168

20

27

03

10

17

24

Figure 7.21 Ozone Diurnal Variation, Winter 197


Piccadilly Bolton Bury Roadside South Day time 10-15 hrs. Night time 18-06 hrs.

30

25

20

ppb

15

10

0
01/01/97 08/01/97 15/01/97 22/01/97 29/01/97 05/02/97 12/02/97 19/02/97 26/02/97

Figure 7.22 Ozone Diurnal Variation, Summer 1997


Piccadilly Bolton Bury Roadside South Day time 08-20 hrs. Night time 22-03 hrs.

60

50

40

ppb

30

20

10

0
97 97 97 97 97 97 97 97 97 97 97 97 97 8/ /0 31 6/ 6/ 6/ 6/ 6/ 7/ 7/ 7/ 7/ 8/ 8/ 8/ /0 /0 /0 /0 /0 /0 /0 /0 /0 /0 /0 /0 /0 8/ 97

In order to understand the two plots above, the concentration of NO has to be analysed in the same context. Figures 7.23 and 7.24 show the diurnal variation of the NO concentration during the same two periods.

01

08

15

22

29

06

13

169

20

27

03

10

17

24

Figure 7.23 NO Diurnal Variation, Winter 1997


Piccadilly Bolton Bury Roadside South

Day time 10-15 hrs. Night time 18-06 hrs.

450 400 350 300

ppb

250 200 150 100 50 0


01/01/97 08/01/97 15/01/97 22/01/97 29/01/97 05/02/97 12/02/97 19/02/97 26/02/97

Figure 7.24 NO Diurnal Variation, Summer 1997


Piccadilly Bolton Bury Roadside South Day time 08-20 hrs. Night time 22-03 hrs.

180 160 140 120

ppb

100 80 60 40 20 0
97 97 97 97 97 97 97 97 97 97 97 97 97 8/ /0 31 6/ 6/ 6/ 6/ 6/ 7/ 7/ 7/ 7/ 8/ 8/ 8/ /0 /0 /0 /0 /0 /0 /0 /0 /0 /0 /0 /0 03 10 17 24 /0 8/ 97

NO2 is responsible for all the photochemical production of O3 and it plays an important role in PM10 formation. The diurnal plots are presented in Figures 7.25 and 7.26

01

08

15

22

29

06

13

170

20

27

Figure 7.25 NO2 Diurnal Variation, Winter 1997


Day time 10-15 hrs. Night time 18-06 hrs.

70

60

50

40

ppb
30 20 10

0
01/01/97 08/01/97 15/01/97 22/01/97 29/01/97 05/02/97 12/02/97 19/02/97 26/02/97

Figure 7.26 NO2 Diurnal Variation, Summer 1997


Day time 08-20 hrs. Night time 22-03 hrs.

ppb

97

97

97

97

97

97

97

97

97

97

97

97

97 8/ /0 31

6/

6/

6/

6/

6/

7/

7/

7/

7/

8/

8/

8/

/0

/0

/0

/0

/0

/0

/0

/0

/0

/0

/0

/0

Temperature is one of the most important meteorological parameters, especially in terms of the diurnal differences that pollutant concentrations can exhibit.

01

08

15

22

29

06

13

171

20

27

03

10

17

24

/0

8/

97

The next two plots present the diurnal differences between day and night and between winter and summer. The variation shown in both figures represent the typical temperature distribution, generally higher during the day and lower in the night. Also typical are the winter-summer distributions: higher in the summer and lower in the winter. Figure 7.27 Temperature Diurnal Average, Winter 1997
Day time 10-15 hrs. Night time 18-06 hrs.

14 12 10 8

Deg. C

6 4 2 0 -2 -4
01/01/97 08/01/97 15/01/97 22/01/97 29/01/97 05/02/97 12/02/97 19/02/97 26/02/97

Figure 7.28 Temperature Diurnal Average, Summer 1997


Day time 08-20 hrs. Night time 22-03 hrs.

30

25

20

Deg. C

15

10

0
97 97 97 97 97 97 97 97 97 97 97 97 97 8/ /0 31 6/ 6/ 6/ 6/ 6/ 7/ 7/ 7/ 7/ 8/ 8/ 8/ /0 /0 /0 /0 /0 /0 /0 /0 /0 /0 /0 /0 01 08 15 22 29 06 13 20 27 03 10 17 24 /0 8/ 97

172

7.4 Statistical Analyses


Statistical analyses were performed to the available data in order to assess the possible associations between PM10, O3 and their precursors together with some of the basic meteorological parameters. Ordinary Least Square regression (OLS) was performed to the relationship between PM10 and SO2, NO2, relative humidity, temperature and wind direction. The coefficient of determination (COD or R2) and the significance level according to the F statistic are presented in Table 7.2. OLS regressions were performed to the data for PM10, NO2, SO2 and the basic meteorological parameters. In the ozone case, regressions were performed to the relationship between O3 and NO, NO2, relative humidity, temperature and wind speed. Multiple regression analysis was also performed for the relationship between PM10 vs. SO2, NO2 and wind speed as these variables proved to be statistically significant according to the coefficient of determination (COD), shown in Table 7.2. Ozone proved to have a relatively strong relationship with NO relative humidity and NO2, thus multiple regression was applied using these parameters. Table 7.3 Statistical Analyses and Durbin-Watson Test Results Pollutant SO2 NO2 Temp. Rel. Hum. Wind Speed Mult. Reg. (SO2, NO2, WSP.) Mult. Reg. (SO2, NO2) NO2 NO Temp. Rel. Hum. Wind Speed Mult. Reg. (NO, NO2, RH) COD (R2) 0.483 0.603 0.006 0.004 0.157 0.652 0.652 0.265 0.398 0.092 0.257 0.097 0.497 Significance (F statistic) P = 0.99 (339.4) P = 0.99 (552.4) P = 0.87 (2.2) 0.85 P = 0.79 (1.6) 0.84 P = 0.99 (67.8) 0.99 P = 0.99 (225.9) P = 0.99 (338.7) P = 0.99 (131.2) P = 0.99 (240.1) P = 0.99 (37.0) 0.58 P = 0.99 (125.8) P = 0.99 (39.1) 0.54 P = 0.99 (119.1) DurbinWatson Test. 1.16 0.99

PM10

1.08 1.05 0.56 0.80 0.85 0.97

O3

The relationship between PM10, ozone and their respective precursors is relatively strong, as suggested by the results presented above. The estimators of the regression analysis may however, be flawed because the residuals are autocorrelated, according to the Durbin-Watson statistic test (Clark and Husking, 1986). On the other hand, the formation conditions can be observed from the relationship plots. Figures 7.29 to 7.33 show the relationship between PM 10 and SO 2 , NO 2 , temperature, relative humidity and wind speed. 173

Figure 7.29 PM10 vs. SO2, 1997


PM10 OLS PM10

90 80 70 60 50 40 30 20 10 0
0 5 10 15 20 25 30

y= 1.8902 x 9.5616 2 R =0.5029

PM10

SO 2

Figure 7.30 PM10 vs. NO2, 1997


PM10 OLS PM10 y= 1.0972 x 1.536 2 R =0.6431

90 80 70 60 50

PM10

40 30 20 10 0
0 10 20 30 40 50 60

NO2

174

Figure 7.31 PM10 vs. Temperature, 1997

90 80 70 60 50 40 30 20 10 0
-5 0 5 10 15 20 25

PM10 OLS PM10 y= 0.1982 x 26.168 2 R =0.0088

PM10

Temp.

Figure 7.32 PM10 vs. Relative Humidity, 1997

90 80 70 60 50

PM10 OLS PM10 y= 1.123 x 14.32 2 R =0.0113

PM10

40 30 20 10 0

20

40

60

80

100

120

RH

175

Figure 7.33 PM10 vs. Wind Speed, 1997


PM10 OLS PM10

90 80 70 60 50

y= 1.0653 x 32.497 2 R =0.1626

PM10

40 30 20 10 0
0 5 10 15 20 25

WSP

Ozone concentration in the troposphere is dependent on the concentration of NO and NO2. The relationships between O3 and these two pollutants and between the basic meteorological parameters are displayed in Figures 7.34 to 7.38. Figure 7.34 Ozone vs. NO, 1997
PM10 OLS PM10

50 45 40 35 30

y= 1.123 x 14.32 2 R =0.0113

O3

25 20 15 10 5 0
0 50 100 150 200 250

NO

176

Figure 7.35 Ozone vs. NO2, 1997


O3 OLS O3

50 45 40 35 30 25

y= 0.4643 x 23.85 2 R =0.2181

O3
20 15 10 5 0
0 10 20 30 40 50 60

NO2

Figure 7.36 Ozone vs. Temperature, 1997

O3 OLS O3

50 45 40 35 30

y= 0.4387 x 8.454 2 R =0.0794

O3

25 20 15 10 5 0
-5 0 5 10 15 20 25

Temp.

177

Figure 7.37 Ozone vs. Relative Humidity, 1997


O3 OLS O3

50 45 40 35 30

y= 0.3751 x 42.755 2 R =0.2009

O3

25 20 15 10 5 0
0 20 40 60 80 100 120

RH

Figure 7.38 Ozone vs. Wind Speed, 1997


O3 OLS O3

50 45 40 35 30

y= 0.572 x 8.5507 2 R =0.0883

O3

25 20 15 10 5 0
0 5 10 15 20 25

WSP

178

7.5 Wind Analyses


All the wind information and in general all the synoptic and climatic data were obtained from the United Kingdom Meteorological Office, in the British Atmospheric Data Centre. The prevailing wind direction for the GMA during 1997 was determined by calculating the frequency of occurrence against eight categories related to the cardinal points. The different bins were: north (338 22), north-east (23 67), east (68 112), south-east (113 157), south (158 202), south-west (203 247), west (248 292), and north-west (293 337). Figure 7.39 shows the prevailing wind rose during 1997 at the South AUN station. Figure 7.39 Prevailing Wind Direction, 1997

Wind frequency

Pollutant roses were prepared in order to examine the possible sources of PM10, O3 and their precursors. Although synoptic data is only available in one of the AUN stations, South, the values are taken to be representative of the whole region, and in the absence of more data, the values are used in all the stations. The pollutant roses were calculated using the same cardinal bins as with the wind frequencies, but they were coupled with the appropriate concentration values of the different pollutants. Three stations were selected for this analysis, South, Piccadilly and Bolton. These three represent the whole area relatively well and follow the trajectory of the urban plume. Figure 7.40 shows the pollutant rose for PM10 in Piccadilly and Bolton. The South station is absent as it does not monitor particulates.

179

Figure 7.40 PM10 vs. Wind Direction, 1997

Picadilly PM10 Bolton PM10

The two precursors of PM10 are monitored in all three stations, hence their graphs are shown in Figures 7.41 and 7.42. Figure 7.41 SO2 vs. Wind Direction, 1997

South SO2 Picadilly SO2 Bolton SO2

180

Figure 7.42 NO2 vs. Wind Direction, 1997

South NO2 Picadilly NO2 Bolton NO2

Ozone is related to NO2 and NO via reactions 3.9, 3.10 and 3.12, the first two in the generation process while the other in the destruction. Figures 7.43 and 7.44 show the roses for O3 and NO respectively. Figure 7.43 O3 vs. Wind Direction, 1997

South O3 Picadilly O3 Bolton O3

181

Figure 7.44 NO vs. Wind Direction, 1997

South NO Picadilly NO Bolton NO

7.6 Conclusion
Time series were calculated to understand the magnitude and distribution of the concentration of PM10 and ozone. Their precursors and destroyer species were also analysed in order to comprehend their relationships and contributions to both the particles and ozone. The meteorological trends had the same purpose; to examine the factors affecting the formation and transformation of the two pollutants. Diurnal distributions were graphed to observe the behaviour of the concentrations during the day and night time and their inter-seasonal fluctuations. Statistical analyses were undertaken to study the relationship between PM10, O3 and their associated pollutants and meteorological parameters. Finally, pollutant and wind roses were prepared to analyse the potential sources of the pollutants. All the work presented here will be placed into perspective in the next chapter, where the behaviour, sources, formation and transformations of PM10 and O3 will be analysed.

182

Discussion of the Results from the GMA


Chapter Eight

8 Introduction
All the information gathered for the analyses in the GMA came from the AUN monitoring stations. These are distributed within the city centre and around the area (shown in Figure 7.1). Thus, they represent the concentrations of urban and suburban locations. Apart from the stations located in Piccadilly and Town Hall, most are located in more remote areas. All the stations have similar sources of the pollutants they monitor with the exception of Bury. This site is strongly affected by motor exhaust pollutants and it reports very high concentrations of PM10, SO2 and NO and a lower value of O3 (due to its local titration by the very high NO concentrations), all relative to the other sites. The seasonal variations are similar to the other sites indicating that the processes are similar but with higher local sources. The following discussion explains the results of GMA and concentrates on the time series of PM10 and ozone. Both pollutants are characterised by their dependence on the concentration of other atmospheric contaminants. As in the previous discussion, time series of PM10, SO2, O3, NO, and NO2 will be discussed and the basic meteorological parameters will be incorporated into the analysis. Particular to the atmosphere of the GMA, is the nature of the PM10, as they seem to be mostly of secondary origin, meaning that they are smaller than 2.5 m in diameter. They follow the trends of SO2 and NO2 very closely. Ozone on the other hand seems to be mostly destroyed by its reaction with NO, a feature readily seen in their respective plots. The diurnal analyses presented in the previous chapter will be used to assess the effect of insolation, meteorology and social activity in the area. Assuming that both PM10 and O3 are affected by photolysis, the amount of solar radiation is critical. The level of activity in the city will be reflected in the concentration of NO (and therefore of NO2 and O3), SO2 and PM10, especially, since they are mainly emitted or formed by automobile exhaust emissions. The natural separation of the schedules assigned to the winter and summer will aid in the analysis of the effects of meteorological parameters, especially temperature. Statistical analyses were performed on the major precursors of ozone and PM10. The results will be interpreted and from the graphs, the conditions leading to their formation will be considered. Finally, the rose diagrams will be used to make an assessment of the possible associations of the two pollutants and their precursors.

8.1 Time Series Analysis


Annual time series of PM10, O3, SO2, NO and NO2 were calculated to explain the influence of meteorology on the formation and transformation of the two former pollutants. Data from five monitoring stations within the GMA were used, the location of which are illustrated in Figure 7.1. 185

The meteorological parameters were obtained from the South station, near Manchester airport in the south of the region. In the following section the influence of temperature, relative humidity, pluvial precipitation and wind speed will be interpreted and explained. The PM10 averages and those of SO2, NO2 and NO were constantly affected by the concentrations reported at Bury, a station next to a motorway with traffic in excess of 200,000 vehicles per day. The averages reported in this station are consistently higher every month throughout the year. Based on this information, it can be stated that all these pollutants are emitted (or produced through oxidation reactions, as in the case of NO2) by automobile exhaust and are therefore all associated by their source. When this station is excluded from the main trends, the values are lowered but the curves retain their shape. Figures 7.2, 7.4, 7.9 and 7.8 display the annual trends of the pollutants just mentioned, and a different curve is included to portray the average trend excluding the values reported at Bury. The concentrations reported by this station will, however, be included in the analysis, as they represent the values from one of the most important sources of these pollutants, the automobile exhaust (de Icaza, Choularton, 1999).

8.11 PM10
The annual trend for PM10 during 1997 shows a relatively small seasonal fluctuation, Figure 7.2. The time series was distinguished by a very high concentration during January, averaging 37g/m3, and a relatively low during February, average 20 g/m3. The reason for this change is centred in the meteorology of the two months. The temperature trend (Figure 7.12) and the weather logs for this period (RMS, 1997a and 1997b) reveal that January was relatively cold, with an average temperature of 3 C, and was dominated by anticyclonic weather patterns coming from the Atlantic. This translated into low mixed layers, lower wind speeds (6 m/s) and higher loadings of PM10 generated by motor vehicles being trapped in the urban area. During the second month however, the weather became milder, the temperature increased to 6 C, and strong westerly winds (13 m/s) coupled with deep mixing dominated the period, normal for cyclonic weather conditions. The particle loadings decreased markedly during February due to vertical mixing and dispersion (de Icaza, Choularton, 1999). During March the concentration of PM10 returned to its average value of around 25 g/m3 and from then on, it followed a normal tendency; lower in the summer, with an average value of 21 g/m3 during July, and slightly higher in the equinox and winter seasons, with average values of 26 g/m3 and 24 g/m3 in October and December respectively. In the time series, PM10 the average values plunged during June and September, explained by the increase in wind speed and the hectic weather conditions experienced during the second half of the June and the first half of September (RMS, 1997c, 1997d).

186

The relationship between higher wind speeds and lower PM10 loadings can be explained through the nature of the aerosol. Smaller particles, emitted mostly by motor exhausts, tend to be trapped in the lower troposphere at times of atmospheric stability; a condition leading to high PM10 concentrations and typically related with lower wind speeds. This is mainly due to the low mixed layers and therefore the lower levels of dilution associated with these conditions. Furthermore, particles affected by this relationship would tend to fall in the smaller size spectra (smaller than 2.5 m in diameter) as the coarse fraction (2.5 - 10 m) would tend to originate from dust and crustal elements and would therefore tend to be resuspended by increasing wind speed. Moreover, in an area like the GMA with relatively constant pluvial precipitation throughout the year, the larger particles, related to road dust and soil, would tend to stay settled in the ground. Overall, the information from the relationship between wind and particle loadings suggest that the majority fall in the smaller range, below 2.5 m in diameter. Harrison (1997) in a study undertaken to study particles in Birmingham UK, found that PM2.5 were negatively correlated with wind speed, indicating that their concentration would decrease with increasing wind speed. On the other hand, coarse particles (between 2.5 and 10 m) tended to have either a positive correlation or no correlation. This is due, as explained, to the fact that the larger particles are associated with dust, soil and other crustal elements that can usually be resuspended by the wind. In dryer environments or in periods without rain, this could translate into higher PM10 concentrations. Here in the GMA, however, high particle loadings usually occurred under low wind speed regimes, indicative of the large proportion of PM2.5.

8.12 O3
The location of the station at Bury affected the O3 concentration. As explained above, it is situated next to a motorway, with high NOx emissions and therefore much lower ozone concentrations (titration through reaction 3.12 in chapter 3), all relative to the rest of the stations (de Icaza, Choularton, 1999). Figure 7.7 shows the monthly averages in each of the stations and it can be seen that the concentrations at Bury were much lower than in the rest of the stations. This difference is especially highlighted during the summer, the period with the highest photochemical potential. The ozone time series shows a marked seasonal variation with a spring maximum, consistent with its photochemical origin and shown in Figure 7.6. It is high during the spring and summer, with average values of 18 ppbv in April, 19 ppbv in May, 18 ppbv in June, 15 ppbv in July and 15 ppbv during August, when photochemistry is at its highest. Also consistent with its photochemical nature, O3 values were low during the autumn and winter seasons, with average loadings of 11 ppbv in September, 8 ppbv in October, 6 ppbv in November and 9 ppbv during December, when insolation is lower.

187

It is noticeable as well that with the exception of Bury, ozone concentrations were highest at the suburban sites away from the strongest primary sources. In general, ozone concentrations increased in a south to north trajectory (Figure 7.7), consistent with the prevailing wind (de Icaza, Choularton, 1999). In the South station the concentration ranged from 7 to 18 ppbv while in Piccadilly the range was from 5 to 22 ppbv and in the northern most station, Bolton, the values ranged from 7 to 25 ppbv. Bury, as expected, reported the lowest O3 concentrations, due to its titration by NO. These values show that the upper limits of the average loadings increased from south to north, suggesting that O3 was formed within the urban plume. Apart from the photochemical aspect of ozone formation, its concentration is affected by the local weather conditions. The low value reported during January and the relatively high concentration in February were related to the weather phenomena occurring during this period. As explained before, anticyclonic conditions dominated the GMA during January, producing high concentrations of PM10, NOx and SO2. At the same time, the high concentrations of NO titrated large quantities of ozone, thus reducing its tropospheric loading. The reversed happened during February: cyclonic weather dominated the area, thus decreasing the NO concentrations and reducing the titration of O3 (de Icaza, Choularton, 1999). Normally, however, in areas where ozone concentrations are more related to its photochemical production, anticyclonic conditions would promote high photochemical activity and therefore large quantities of O3. Under stable conditions, NO2 and VOCs are able to interact and in combination with clear skies promoting insolation, produce larger loadings of ozone. It is only in urban areas like the GMA, that the former situation occurs, where anticyclonic conditions prevent high ozone loadings through its titration with NO.

8.13 Meteorology
The concentration of pollutants is highly dependent on the meteorological conditions of the atmosphere. The synoptic data studied here were obtained from a single station in the south of the region, but it is generally believed that the values are representative of the whole area. The main parameters analysed were temperature, relative humidity and wind speed. Temperature is important as it influences the formation of pollutants and because it can be used as a proxy for estimating the height of the mixed layer. As a rough simplification, when the temperature is relatively high, the air expands and thermals from the ground raise the mixed layer to higher heights. The temperature time series shows a relatively standard distribution, low during the winter, higher in the equinoxes and a climax in the summer. The values fluctuated between 3 and 18 degrees Celsius. The influence of temperature on the formation and transformation of O3 and PM10 can be seen when their concentration values during the first two months of the year are compared (Figures 7.6 and 7.2).

188

When the temperature was low, the concentration of PM10 increased (due to low mixed layers trapping the particles close to the ground) while O3 decreased (owing to the entrapment of NO). This is partly due to low thermal inversions experienced during January, which tend to trap the pollutants near the surface. The big difference in concentration observed during first two months were also related to the occurrence of anticyclonic conditions throughout January and the very hectic cyclonic weather during February (de Icaza, Choularton, 1999). The former favouring thermal inversions (and usually a low mixed layer during cold winters), while the later tends to enhance turbulence and dispersion. The temperature distribution during 1997 can be observed in Figure 7.12. Relative humidity is a measure of the amount of moisture in the air and as such, it influences the formation and removal of pollutants. The trends of PM10 and of O3 were not particularly affected by the changes in relative humidity, as this remained relatively constant throughout the year. The time series of relative humidity during 1997 is shown in Figure 7.13. Windspeed is the ultimate parameter driving mechanical turbulence and therefore the level of dispersion within the troposphere of an airshed. The wind profile of 1997 is shown in Figure 7.14 and it can be used to explain some of the irregularities of the trends followed by O3 and PM10. The year started with a relatively low average wind speed (6 metres per second (m/s)) during January (related to the anticyclonic conditions, causing high atmospheric stability) and then it rose to around 13 m/s in February. As it is shown in Figures 7.2, the average values of PM10 plunged between these two months. Figure 7.8 also shows a relatively important reduction in the concentration of NO between these two months, explaining the increase in O3, Figure 7.6. Furthermore, when the windspeed increased again during June, the trends of O3 and PM10 were affected in the same manner; the former increased as the NO concentration decreased and the later was lower. During the rest of the year, the average wind speed fluctuated from around 6m/s to 9 m/s, with a maxima in June, September and December, and a minima in May, July and October.

8.2 Pollutant Associations


Primary and secondary atmospheric pollutants differ in their origin. The former are directly emitted, while the later are formed through reactions with other pollutants in the atmosphere. Ozone is secondary, while the particulates can be either one. Secondary PM10 depends on the concentration of SO2 and NO2, both contributing to the photochemical formation of particulates through their interaction with the hydroxyl radical (OH) and through their aqueous phase reactions in the atmosphere. Tropospheric O3 depends on the photolysis of NO2 while at the same time its concentration is dictated by its titration with NO.

189

The associations between primary and secondary pollutants depend on many factors but probably one of the most important is the concentration of precursor species. In the following sections, the interactions between PM10, O3 and their precursors will be examined and the influence of temperature and windspeed will be interpreted.

8.21 PM10 Associations


Aerosols are a very peculiar contaminant as they can be directly emitted into the atmosphere as primary particles or they can form by reactions of different atmospheric species and become secondary. The principal precursors of PM10 are SO2 and NO2, through both gas and aqueous phase reactions (the formation mechanisms are outlined in chapter two, section 2.23). Sulphur dioxide can be oxidised photochemically by reacting with OH to produce sulphuric acid (H2SO4), reactions 2.5 through 2.7. In the aqueous phase, several atmospheric species may catalyse its oxidation to form the sulphate aerosol, reactions 2.10 to 2.13. Figure 7.4 shows the time series of SO2 during 1997, a trend closely followed by PM10. The concentration is high during January and low during February, for the same reasons as the particulate trend. When the curve excluding Bury is compared with the PM10 trend, the similarities increase. Apart from the first two months (average SO2 concentrations during January and February were 10 ppbv and 5 ppbv respectively), both pollutants return to their average values during March, (SO2 was around 6 ppbv), and they both reported low concentrations in June (4 ppbv). As in the case of the PM10, the lower SO2 value in June was also related to the higher wind speeds and the hectic weather conditions. In the same manner as sulphur dioxide, NO2 may react with OH to produce nitric acid (HNO3) via reaction 2.8. In the aqueous phase, it may oxidise by reacting with HSO3- or O3, also to ultimately form HNO3, reactions 2.13 and 2.14 to 2.16. Figure 7.9 shows the NO 2 time series. This trend is also very similar to the particulate average values. Once again, when the curve excluding Bury is compared with the PM10 trend, the similarities increase. The first two months reflect the influence of the weather with concentrations during January and February in the order of 29 ppbv and 19 ppbv respectively. During March, the NO2 values returns to their normal tendency with an average value of 23 ppbv. The curve excluding Bury shows lower NO2 values during June and September, average concentrations of 17 ppbv and 18 ppbv, mainly for the same reasons as with the particles and SO2, higher wind speed and more hectic weather patterns. In the time series, it can be observed that the concentration of PM10 remained relatively constant and up to a certain extent decreased throughout the periods of highest photochemistry, April to July. These results suggest that the sources of PM10 are similar throughout the year, indicating that photochemistry plays little role in their formation, that is, through the conversion of SO2 and NO2 into H2SO4 and HNO3 via reaction with OH. 190

Based on the similarity of the time series of PM10, SO2 and NO2, and on the relationship between the particles and wind speed, it can be speculated that most of the particulates were secondary in nature (less than 2.5 m in diameter) and that NO2 had a larger contribution to the formation of PM10 than SO2. Further evidence emanates from size distribution and chemical composition data of the aerosol in the GMA. Inglis (1998) reported data sampled throughout the summer and autumn of 1997 and the results showed that PM2.5 represented 74% of the total PM10 during the whole sampling period. The breakdown by month showed a range from 76% to 83% in different months of the sampling period, with the highest proportion occurring during November. The chemical composition of the aerosol sampled is displayed in Table 8.1. The values show that most of the PM10 were associated with SO2 and NO2 and that ammonium (NH4) had an important contribution to the formation of the secondary particles. During the summer when NH4 was at its highest, the contribution of the nitrates became more important, and the sulphates undertook a secondary role in the formation of the aerosol. This is consistent with the theory as NH4 will react preferentially with SO2 and only when there is enough, will it also react with NO2. The details of the reaction mechanisms and of the role of ammonium in the formation of nitrate and sulphate aerosols are given in section 2.23 of chapter two. Although organic carbon was not measured by Inglis (1998), Harrison (1997) studied the sources and processes affecting PM10 and PM2.5 in Birmingham, UK. In his study, he calculated that 20% of the PM10 were composed of organic carbon, 18% of elemental carbon and the rest of the ionic species accounted for 29%, while the remaining compounds were attributed to unknown or not measured. Although Birmingham and the GMA are relatively different areas, the sources of PM10 would tend to be similar as they both have large contributions from motor exhaust (seen in the large quantities of NO loadings, Birmingham reported a mean NO concentration of 34 ppb (Harrison, 1997), while the average concentration in the GMA was 42 ppb). Table 8.1 Chemical Composition of PM 10 in the GMA Period Spring 1996 Summer 1996 Summer 1997 Winter 1997 SO4 39.14% 33.70% 26.75% 27.99% NO3 29.27% 22.26% 35.78% 24.61% NH4 20.06% 20.53% 29.01% 20.86% Other species 11.53% 23.51% 8.46% 26.54%

191

The results reported by Inglis (1998) show that sulphates contributed the larget proportion of the aerosol mass and that nitrates had a secondary role. The similarity of the time series of PM10, SO2 and NO2 suggest otherwise, that NO2 was more related to the particles than SO2. A possible explanation is that NOx and PM10 were both emitted, to a large extent, by the same source and therefore had relatively more similar trends. Furthermore, NO2 oxidises to form the nitrate aerosol under the presence of ammonia, a molecule that in the GMA tends to have relatively large loadings, in the order of 4 5 ppb (Choularton, 1999). This evidence therefore suggests that the contribution of the nitrate species was relatively more important to particle formation than sulphate. Generally through the year, there was very little variation in PM10 except that the concentrations were highest during stagnant periods in winter with a low level temperature inversion and lowest during convective days in the summer (de Icaza, Choularton, 1999). This evidence further corroborates the secondary origin of the particles, as the smaller particles will be associated with the above precursors and emitted by motor exhaust while the larger particles would tend to be associated with dust and other crustal elements. At periods of lower wind speeds, related to high-pressure systems, the concentration of the secondary particles would tend to increase due to deeper mixing and lower levels of dilution in the atmosphere.

8.22 O3 Associations
Ozone is a secondary pollutant produced by the photolysis of NO2 in the presence of hydrocarbons. The concentration of O3 is therefore tightly linked to the concentration of NOx (NO + NO2) and their reaction systems. The photolysis of nitrogen dioxide and the ozone formation mechanisms are detailed in section 3.31 and throughout chapter three. Ozone concentrations were very low during January with an average concentration of 5 ppbv, lower than typical background levels of 10 to 20 ppbv. Two reasons account for this relatively low loading, the scarce amount of insolation, reducing the photochemical potential and the very large concentrations of NO. As explained before, the concentration of NO during this month was extremely high, in the order of 98 ppbv, mainly because of the weather patterns experienced during this month. As the concentration of NO increased, large quantities of O3 were titrated, thus reducing its tropospheric concentration; the relationship can be seen in Figure 7.10. During February ozone concentrations increased due mainly to milder weather, higher mixing (reducing the NO concentration) and the onset of photochemistry (de Icaza, Choularton, 1999). Tropospheric ozone in the GMA is generally depleted rather than produced. This can be readily seen in Figure 7.10.

192

The curve representing ozone matches very closely, in the opposite direction, the NO trend. As stated before, when the concentration of NO was high, the respective O3 concentration was low. This is true in all the months except for June, when the NO concentration was lower (15 ppbv) than O3 (18 ppbv). This trend was only related to the NO values reported at Bury, as the trend excluding this station showed no deviation from the normal relationship, with a NO concentration during June equalled to 11 ppbv and an ozone concentration equalled to 20 ppbv, Figure 7.11. In urban areas over the UK, ozone depletion via reaction with NO is greater on average than the photochemical production (PORG, 1997). Under these type of conditions, NO controls the concentration of O 3 and not the photochemical production through NO2. The forward reaction, that of the formation of O3 through the photolysis of NO2 (reaction 3.12), is clearly represented in their respective trends. Under normal conditions, large quantities of NO2, in the presence of sunlight and hydrocarbons, would produce large quantities of O3, but in the formation process, the precursor species is depleted. The overall result is that high concentrations of ozone translate into low NO2 loadings. The relationship is therefore similar to the NO and O3 relationship, following opposite trends. The curves in Figure 7.10 display this scenario, low O3 concentrations are equalled by high NO2 and vice versa, i.e. during January ozone was 5 ppbv while the NO2 concentration totalled 24 ppbv, during June both pollutants were 18 ppbv and during November, ozone was 6 ppbv and NO2 was 25 ppb. The concentration of hydrocarbons, or more specifically volatile organic compounds (VOCs), in relation to the concentration of NOx species either promotes or hinders the production of O3. In urban areas where NOx are prevalent, like the GMA, the limiting factor affecting such production is the availability of VOCs, a condition generally referred to as VOC-limited. In a map of the whole UK, the GMA was determined to be VOC limited, and a reduction of O3 could be achieved only by a reduction in the VOC emissions (PORG, 1997).

8.3 Diurnal Analyses


Graphs showing the different pollutant distributions during the day and night and during the summer and winter are important because they display the important factors contributing to their concentration in the atmosphere. As in the past analyses, the purpose is to determine the formation and transformation mechanisms of O3 and PM10. Studying their diurnal behaviour and that of their precursors can provide information on the processes influencing their concentration in the GMA. Like before, PM10 will be related to SO2 and NO2 because they seem to contribute to their formation, especially through their aqueous phase reactions.

193

Ozone will be associated with NO, mainly because it is the species controlling its concentration and to NO2 as it is generated in the urban plume. The diurnal trend of PM10 is fairly standard with mostly higher concentrations in the day. During the fist part of the year, the concentrations were generally higher than in the more convective summer, consistent with the lower temperatures and therefore lower mixed layers, shown in Figures 7.17 and 7.18. The diurnal distribution of SO2 was very similar, mainly higher during the day and lower in the summer, Figures 7.19 and 7.20. NO2 also showed a typical diurnal distribution, mostly higher in the day and during the summer, Figures 7.25 and 7.26 for the winter and summer respectively. The distributions of the three pollutants showed that they are mostly formed during the daytime and that the concentrations were higher in the colder and more atmospherically stable months. The loadings of the three pollutants were generally higher in January than in February but in both months, the distribution between night and day was relatively standard, that is, higher in the day. The larger loadings during the day reflect the influence of social activity as the precursor species were mainly produced or emitted during this time. Ozone on the other hand shows a more complicated diurnal trend. In many days, the concentration was lower in the daytime than the respective nights, contrary to its photochemical origin, shown in Figure 7.21 for winter and Figure 7.22 for the summer. Ozone is normally titrated by NO during both day and night, but NO is mainly emitted during the day by motor exhausts, as can be seen in Figures 7.23 and 7.24. The result is that there is more NO during the day time to titrate O3 and thus reduce its tropospheric concentration. Ozone, however, is only produced during the daylight hours by the photolysis of NO2 and the higher concentrations during the night indicate that it must have originated elsewhere. This fact further corroborates the hypothesis that O3 was produced in the urban plume and transported into the GMA from the adjacent rural areas. This can be readily seen in the Figure 7.7, where all the stations are shown and the ozone concentration increased from south to north, in accordance with the prevailing wind of the area. The diurnal distribution of temperature was relatively standard, mostly lower during the night and higher in the summer, Figures 7.27 and 7.28. Diurnal differences were much higher in the summer, 4 C on average, than in the winter, 1.6 C. A relatively steep progression of higher temperatures can be observed in the winter plot, reflecting the radical change in the weather patterns. The influence of the diurnal and inter-seasonal differences of temperature on the formation of PM10 and O3 has been highlighted throughout the chapter. When low temperatures are coupled with anticyclonic conditions, dispersion is hampered and the mixed layer is low, ideal for higher pollutant loadings. This is true for most pollutants, but not for O3, as its concentration is dependent on the NO loading.

194

On the opposite end, higher temperatures translate into higher mixed layers that when coupled with more hectic weather patterns, and usually favour lower atmospheric concentrations of pollutants. Again, ozone is the exception with higher concentrations during convective periods. The normal pattern of diurnal fluctuations is to observe higher temperatures during the day and the inter-seasonal temperature differences to be higher throughout the summer, seen in both diurnal plots, Figures 7.27 and 7.28. When these are compared with the pollutant plots, it can be seen that most had higher loadings during the winter and lower during the night. The later situation, however, was related to the lower emissions and formation rates and not to the temperature structure. Diurnal analyses are also useful to relate the concentration of secondary pollutants to the levels of insolation. In the case of PM10, solar radiation is not affecting their formation directly, as their concentration is similar throughout the year, reflecting that photochemistry plays a minor role in their formation. Ozone, on the other hand, can only be created in the presence of sunlight, NO2 and hydrocarbons. The diurnal trends of O3 show this relationship, having larger loadings during the summer, when solar radiation and photochemistry climax.

8.4 Statistical Analyses


In order to study the association between pollutants in a more quantitative manner, statistical analyses were performed. Specifically, Ordinary Least Square (OLS) regressions were applied to PM 10 , O 3 , their precursors and meteorological parameters. The results, included in Table 7.2 of the previous chapter, indicated that PM10 was well correlated with NO2 and SO2, with Coefficients of determination (R2) of 0.60 and 0.48 respectively. Particles were also better correlated with windspeed (R2 = to 0.16) than with temperature (R2 = to 0.006) and relative humidity (R2 = 0.004). Based on these results, multiple linear regression was applied to NO2, SO2 and windspeed and NO2 and SO2, yielding R2s of 0.65 in both analyses. Ozone was well correlated with NO, NO2 and relative humidity, with R2s equal to 0.40, 0.27 and 0.26 respectively. Multiple regression was applied to NO, NO2 and relative humidity for the same reason as the PM10; they were well correlated on their own. The result yielded an R2 of 0.50. The significance levels of all these results were very high, with F values mostly around 0.99 probability, also shown in Table 7.2. The problem with the significance level of these results is that according to the Durbin-Watson test, they cannot be trusted. Consistent with this test, the residuals are autocorrelated and they are not statistically independent. This translates into inefficient estimators and invalidates the t and F statistic test. Nonetheless, the correlations are correct although they cannot be tested for significance. The results of the statistical analyses corroborate the previous hypothesis, that PM10 is being formed by the reactions of NO2 and SO2 in the atmosphere and that wind speed is relevant to their atmospheric concentration.

195

Ozone on the other hand is better correlated with NO than with NO2, a result suggesting its importance on the tropospheric O3 concentration. Apart from the quantitative analyses, the plots can be used to relate the formation conditions of both PM10 and O3. By inspecting Figures 7.29 to 7.33, it can be observed that most PM10 were formed under two temperature regimes, 6C11C and 13C19C. Relative humidity was around 70% 90%, and windspeed averaged 4 8 metres per second. In terms of precursor species, the particles were formed under 2 11 ppbv of SO2 and 12 32 ppbv of NO2. The highest PM10 concentration occurred when NO2 was 52 ppbv, SO2 equalled 21 ppbv and the temperature, relative humidity and windspeed were 2 C, 79 % and 4 m / s. Figures 7.34 to 7.38 show that ozone had a similar picture, being formed under the whole temperature spectrum (9 C 22C), with a relative humidity in the range of 80% 90% and windspeed in the order of 2 10 metres per second. The concentration of NO2 was around 15 35 ppbv and it was titrated at 5ppbv40ppbv of NO. The highest O3 concentration occurred under 9 ppbv NO2, 4 ppbv of NO and the meteorological parameters were; a temperature of 8 C, 67% relative humidity and 14 m / s windspeed. Although these values are very subjective, they do support the hypotheses posed before, O3 concentrations is predominantly controlled by NO and the particle loading is dependent on the concentration of their precursor species.

8.5 Wind Analyses


The purpose of the pollutant and wind roses is to attempt an assessment of the interactions between the pollutants. Once again PM10 will be related to SO2 and NO2, while ozone will be compared with NO and NO2. To increase the significance of the analysis, the urban plume will be followed from the southern most station, South, to Piccadilly in the centre and to Bolton, excluding Bury, the northern most (see Figure 7.1 for a reference on the GMA grid). Bury was not selected as it has a very strong local source and would portray a biased analysis. These stations were chosen because they represent the overall area relatively well and because they follow the trajectory of the prevailing wind. Based on the frequency of occurrence during 1997, Figure 7.39 shows that the prevailing wind came from the south. Figures 7.40 to 7.42 display the direction from where the different concentrations of PM10, SO2 and NO2 originated. These figures portray that PM10 concentrations were quite spread, and that the highest loadings came from the same place as the SO2 and NO2, predominantly from the south-east and east. The pollutant roses corroborated the hypothesis that PM10 were secondary in nature, and that both of its precursors influenced their formation. Figures 7.43 and 7.42 on the other hand, show that O3 was not formed locally, as its precursor concentrations were relatively low in the direction of its highest loadings. NO2 concentrations were largest from the south-east and north, while ozone loadings were higher from the western side of the ordinal points.

196

It can also be observed that the O3 concentrations increased as the urban plume moved from south to north, higher in Bolton than in Piccadilly and even lower in South. Moreover, the highest NO concentrations originated in the north (Figure 7.44) while the O3 rose showed its lowest concentrations coming from that direction, consistent with its titration via reaction 3.12. The conclusions of this analysis corroborate the previous hypotheses, that PM10 was mostly secondary, and that it was mainly formed through the reactions of both NO2 and SO2. Furthermore, ozone was not produced locally but in the trajectory of the urban plume.

8.6 Conclusion
The preceding discussion explains the results of the GMA analyses. Time series were examined to understand the distribution of the pollutant concentration throughout the year. Furthermore, the series were used to compare and assess the influence that precursors had on secondary pollutants. Diurnal and inter-seasonal differences were examined to interpret the influence of social activity, temperature and insolation on the formation and transformation mechanisms of PM10 and O3. In an effort to quantify the relationships, OLS regressions were performed to the predominant associations. To complete the analysis, pollutant roses in the trajectory of the urban plume were examined. The results show that PM10 were secondary in nature and that they were formed by the aqueous reactions of NO 2 predominantly, and SO 2 . Photochemical formation of particles was proven to be minor. It was also shown that temperature coupled with weather patterns affect the concentration of PM10, especially since both of its precursors were severely affected as well. The diurnal distribution showed that particles had higher loadings during the day, a reflection of social activity, as their precursors were also higher. The statistical analyses provided evidence that NO2 was the predominant contributor to PM10 formation, followed by SO2. The statistical analyses also showed that wind speed was important but temperature and relative humidity were not well correlated with the particles. The pollutant roses corroborated the hypothesis proposed, that PM10 are secondary and that NO 2 and SO 2 contributed to their formation. The analyses corresponding to O3 showed that most ozone was produced in the urban plume, away from the strongest local sources of NO. Titration rather than photochemical production controlled the tropospheric ozone concentration. It was also shown that temperature and weather conditions affected the O3 concentration, but very differently to the rest of the pollutants. During stagnant periods ozone decreased, while in more hectic conditions, it increased.

197

Diurnal distributions showed that O3 was sometimes higher during the night, contrary to its photochemical origin, but consistent with the hypothesis that it was produced in the urban plume and transported into the GMA from elsewhere. Statistical regression proved that O3 was more affected by NO than by NO2 and that temperature and wind speed did not influence its photochemical formation. The pollutant roses corroborated the generation of O3 within the urban plume, as its concentration increased in a south to north trajectory, the same direction as the prevailing wind. When all the results are analysed together, they yield several conclusions. Atmospheric particles were mostly secondary and were formed during the day through the aqueous phase reactions of NO2 and SO2. Ozone was not produced locally but in the urban plume. Furthermore, NO instead of NO2 controlled its concentration. Meteorological parameters were important, especially temperature and wind speed. Together they determined the dispersion and dilution capacity of the atmosphere and therefore, the formation, transformation and concentration of both ozone and PM10.

198

Comparative Analyisis
Chapter Nine

9 Introduction
In the following sections, the formation and transformation mechanisms of PM 1 0 and O 3 in the MCMA and the GMA will be compared. The basic meteorological parameters will be used to illustrate the influence that temperature, relative humidity, pluvial precipitation and wind speed had on the distribution and concentration of the two pollutants. The nature and associations of PM10 will be discussed in terms of the size distributions and chemical composition. Further associations between the particles and their precursors will be examined through the similarities of their time series and the regression analyses performed. A comparison between the reported data and the conclusions of the analyses reported in the discussion of the results of both locations will help elucidate the true composition of the particles. The formation and removal of ozone from the troposphere of the MCMA and the GMA will be examined and the factors controlling both mechanisms will be compared. In the end, all the information will be placed into context and the different formation and transformation mechanisms of PM10 and O3 in the GMA and the MCMA will be explained.

9.1 Meteorology
In the analysis of the two previous discussions, several conclusions were drawn. One of the most important was the influence of meteorological parameters on the concentration distribution of both PM10 and O3. The analyses undertaken for the MCMA revealed that temperature, relative humidity, pluvial precipitation and under the higher regime, wind speed, affected the concentration of both pollutants. In the GMA, however, relative humidity did not influence the formation or transformation mechanisms of either pollutant while the influence of wind speed and temperature were paramount on the concentration of both PM10 and O3. Figure 9.1 to 9.3 display the average distribution of the four relevant parameters in both agglomerations. The averages of the MCMA were calculated for the three years analysed (1994 1996), while the GMA means represent the concentration during 1997. In all the graphs, the averages are considered to be representative of the periods analysed with the exception of wind speed. The average wind speed in the MCMA fluctuated severely between 1994 and 1995 and therefore this parameter will be graphed with the values of each year separately.

201

Figure 9.1 Temperature Average in the MCMA and the GMA


Temp. MCMA, 1994 - 96 Temp. GMA, 1994 - 96

25

20

Degrees Celsius

15

10

0 Jan Feb Mar Apr May Jun Jul Ago Sep Oct Nov Dec

Figure 9.1 shows that the temperature distributions of the two areas were very different. In the MCMA, being closer to the equator, the temperature remained relatively constant throughout the whole year, with an average fluctuation of 4 C. In the GMA, however, the difference was much higher, with an average variation of 15 C, representative of its location in the northern latitudes. The impact that temperature had on the loadings of both pollutants was more related to the indirect effects that temperature had on the overall troposphere of both areas. As explained in the previous discussions, temperature may be used as a proxy for mixed layer height, with higher temperatures representing higher mixed layers. In general, when the mixed layer is higher, pollutants may disperse more freely and their concentrations are affected through higher dilution factors. When temperature is coupled with atmospheric stability data, more complex relationships can be assessed. A good example would be the difference in concentration of PM10 and other primary pollutants exhibited during the first two months studied in the GMA. During January, the temperature was low and a high-pressure system dominated the area, producing low mixed layers and the trapping of particles and primary pollutants close to the ground. In February, however, the temperature increased and a low-pressure system became dominant, reducing the loadings of most pollutants.

202

Although this relationship is a simplification of the overall process, as wind speed and other factors need to be considered, it exemplifies the impact that temperature had on the overall loading of the particles and other primary pollutants within the atmosphere of the GMA. In the MCMA, temperature fluctuations were relatively minor and the impact can be seen in the seasonal variations exhibited in the pollutant loadings, generally lower during the spring and summer, at the time of the highest temperatures. Figure 9.2 displays the relative humidity trends for both areas, again the values for the MCMA are averages of the three years, while the means for the GMA represent the values for the year analysed. Figure 10.2 Relative Humidity and Rain Averages for the MCMA and the GMA
100 90 80 70 60 100 180 160 140 120

Percentage

50 40 30 20 10 0 Jan Feb Mar Apr May Jun Jul Ago Sep Oct Nov Dec
RH MCMA, 1994-96 RH GMA, 1997 Rain MCMA, 1994-96 Rain GMA, 1997

80 60 40 20 0

As can be observed, the average humidity in the GMA did not fluctuate throughout the year, while in the MCMA there was a definite seasonal variation. The trends exhibited by both areas translated into different formation and transformation mechanisms of both pollutants. In the MCMA, the seasonally distributed relative humidity coincided with the rainy season, which in turn translated into lower loadings of both PM10 and O3 during the summer season. On the other hand, the GMA had a relatively constant humidity and this translated into having different removal processes, which in general were not directly affected by relative humidity or rain. The removal mechanisms of both pollutants were explained in their respective chapters and the influence of rain and relative humidity on the pollutant concentrations in the MCMA were analysed in the discussion chapter of this area.

203

For the purpose of completeness however, the impact that relative humidity and rain had on the removal mechanisms of both pollutants in the MCMA will be discussed. Rain may only develop under high relative humidity conditions. As precipitation developed, the larger coarse particles were removed via impaction with the falling rain, while the presence of water in the surface retarded or prevented resuspension of dust and other crustal elements. The smaller secondary particles, on the other hand, may undergo chemical transformations and become incorporated into the cloud water. Ozone was removed through several mechanisms, including its assimilation into rainwater. This occurred through its reaction with NO2 to produce NO3, which in turn reacted further with another molecule of NO 2 to produce N 2 O 5 . This molecule may undergo hydrolysis to produce nitric acid and become incorporated into rainwater. Further removal occurs via the addition of HOx radicals into cloud water, producing an oxygen ion that may react with O 3 (PORG, 1997). Other important removal mechanisms include photochemical reactions of both O3 and its precursor species, NO2. Reaction 3.25 removes NO2 from the troposphere, while reaction 3.62 removes O3 directly. In both cases the reference species react with HOx radicals and produce nitric acid in the former case, while in the latter ozone is transformed into O2 and a further hydroxyl radical. Although this is a relative simplification of the overall process, it suffices the purpose as the detailed mechanisms were presented in the discussion of the results of the MCMA. The wind speed profiles for the years analysed in the MCMA and in the GMA are displayed in Figure 9.3. Among other factors, the trends show that in the MCMA there was a great difference in the average wind speed reported in 1994, part of 1995 and 1996. During the first year and the first quarter of 1995 el nio phenomenon affected the atmosphere of the MCMA, producing mainly higher wind speeds, on the average 3 5 m/s or 2 to 3 times higher than the norm. This in turn affected the concentration of both PM10 and O3 in the troposphere of the valley. In general, the particles had lower loadings and more ozone was produced under effect of el nio. Generally however, under the higher regime, both pollutants showed a negative relationship with wind speed, showing higher loadings with decreasing wind speed. In the rest of the period, under lower wind speeds (average 1 2 m/s), the concentrations of both pollutant were not affected by changes in wind speed, showing a threshold value in the relationship between pollutant loading and wind speed. As can be seen in the next figure, the average wind speed in the GMA fluctuated vigorously throughout the year.

204

It was very low in January and it increased to reach its highest speed during February. During these two months the weather patters of the area changed dramatically, with a very stable atmosphere during the first and a very hectic situation in the second month. Figure 9.3 Wind Speed in the MCMA and the GMA
14 12 10 8

m/s

6 4 2 0 Jan Feb Mar Apr May Jun Jul Ago Sep Oct Nov Dec

WSP MCMA, 1994

WSP GMA, 1995

WSP MCMA, 1996

WSP GMA, 1997

This change also affected the concentration of both PM10 and O3 in the area. In the former condition, lower wind speed coupled with a relatively low temperature and an anticyclone dominating the area produced very high loadings of PM10, NO, NO2 and SO2, while much lower concentrations of O3 were recorded. As was explained before in the discussion of the results of the GMA, the concentration of O3 was dictated by the loadings of NO, thus the high NO concentration during the first month translated into very low O3 concentrations. The rest of the pollutants were mainly emitted by motor exhaust and were therefore trapped close to the ground by the relatively low mixed layer produced by the low temperature and the very stable air. In the second month, the average wind speed and temperature increased and a cyclonic condition dominated the area. The result was the opposite, generating low PM10, NO, NO2 and SO2 loadings and a much higher O3 concentration. The same relationship was observed, to a lesser extent, during June and September when the wind speed increased and anticyclonic conditions dominated the area.

205

Taken together, the four meteorological parameters, it can be stated that in the MCMA the concentration of PM10 and O3 were mostly affected by temperature, relative humidity, pluvial precipitation and under the higher regimen by wind speed. In the GMA, however, temperature and wind speed dominated the meteorological effects on the loading of both PM 1 0 and O 3 . Relative humidity and pluvial precipitation were relatively constant throughout the year and their impact was not obvious in the pollutant loadings.

9.2 Pollutant Associations


One of the most striking differences between both pollutants in the two different agglomerations was their difference in concentration loadings. In the MCMA, the monthly concentrations of PM10 and O3 were three to nine times higher, while the overall averages were more than twice for the particles and more than three for ozone (PM10 averages were 55 and 24 u/m3 in the MCMA and the GMA, while O3 loadings were 41 and 13 ppbv for the same two agglomerations).

9.21 PM10
Figure 9.4 displays the distribution of the PM10 averages in both urban areas; as before the averages for the MCMA represent the mean concentration for the three years, while in the GMA they were calculated for the year analysed. Figure 9.4 PM10 Average in the MCMA and the GMA
PM10. MCMA, 1994 - 96 PM10. GMA, 1997

80 70 60 50

ug/m3

40 30 20 10 0 Jan Feb Mar Apr May Jun Jul Ago Sep Oct Nov Dec

206

Apart form the obvious loading differences, there were many other factors that distinguish the aerosol in the two agglomerations. Among the most important were their origin. From size distributions, it was found that in the MCMA, around 20% (Miranda, 1994) were secondary (meaning that they were less than 2.5 m in diameter), while in the GMA the percentage was in the order of 75% (Inglis, 1998). This difference is rather significant as the smaller particles have distinct characteristics that make them more hazardous to the environment and to human health. In the smaller range, particles interfere with the visible range of the irradiation spectrum, thus visibility is impaired, while they also settle into deeper parts of the humans lungs (DDF, 1996). Moreover, the smaller size range of particles is associated with SO2 and NO2 through aqueous phase and photochemical reactions, the formation and transformation processes of which were explained in chapter two. The chemical composition of the particles were studied in both agglomerations and in general suggest that the sulphate and nitrate ions were responsible for a large proportion of the formation of secondary aerosol in both places. Table 9.1 displays the average composition of the PM2.5 in the MCMA and in the GMA. As can be observed, in the MCMA the contribution of sulphate takes a secondary role, while in the GMA it was the major component. Moreover, organic carbon represented a large proportion of the mass in the MCMA, while in the GMA, Harrison (1997) reported that, in Birmingham (an area with similar sources of emissions), 37% of the PM10 were accounted for by carbon compounds, 18% elemental carbon and around 20% organic carbon. Table 9.4 Chemical Composition of PM2.5 in the MCMA and the GMA Urban Area MCMA GMA *Not Available SO4 15% 32% NO3 16% 28% OC NH4 N/A* Other Compounds 49% 18%

20% N/A* 23%

(adapted from DDF, 1997 and Inglis, 1998)

The results from the time series analysis suggested the same type of contribution to the MCMA aerosol, with a higher proportion of nitrates, while the sulphates were not well related, as the trend of SO2 was very different from the PM10 series. In the GMA, however, the same result was suggested by the time series, that is, a higher contribution from the nitrate and a smaller proportion from the sulphate species.

207

The discrepancy between the two analyses could have originated from the sources of the pollutants as both are emitted by automobile exhausts and are therefore monitored in the same air mass. The time series for SO2, NO2 and PM10 in the MCMA are shown in Figure 5.21 of chapter five, while the same pollutants in the GMA are displayed in Figure 7.5 of chapter seven. The statistical analyses performed also suggested higher contributions from NO2 to the particles, with a R2 equalled to 0.60 and 0.31 in the GMA and the MCMA respectively, while the results for SO2 yielded lower values, R2 equalled to 0.48 and 0.059. Apart from the time series and the statistical analyses, evidence of the dominant role of nitrate was established by the relatively large loadings of NH3 (4 5 ppbv, Choularton, 1999) in the atmosphere of the GMA. As it was explained in the discussion of the results of the GMA and throughout chapter two, ammonia will normally react with the sulphate ion, but when it is available in large quantities, it will also react with nitrate to fix it onto nitric acid and the aerosol phase. The size distribution of the particles has other repercussions as they behave differently in the atmosphere. Smaller particles show a negative correlation with wind speed while the larger sizes tend to exhibit a positive relationship with increasing wind speed (Harrison, 1997). The relation is explained through their nature, as the larger particles would tend to originate from dust or crustal elements that may be resuspended by the wind and would therefore have increased concentrations with increasing wind speed. The smaller range, on the other hand, originates mostly from combustion sources, mainly motor exhausts, and would tend to disperse and dilute with increasing wind speed. The difference in the type of particles found in both the MCMA and the GMA could be related to the inherent characteristics of their environments. In the MCMA, the climate is usually characterised as semi-arid, although it has a great amount of precipitation. The fact is that the rainy season is very well defined and it predominantly occurs during the summer while the rest of the year is relatively dry, as can be seen in Figure 9.2. Figure 9.4 shows that the average trend of the particles in the MCMA was lower during the periods of high rainfall, indicating that the loadings were suppressed by the falling rain and the wet surface. In the dryer periods, on the other hand, larger particles were resuspended, creating higher loadings of PM10. In the GMA, however, rain is distributed throughout the year and the resuspension of dust occurs at a smaller scale. Although both agglomerations had very large sources of secondary particles and their precursors (as they both have large quantities of motor vehicles and other polluting industries), the dryer environment in the MCMA induced higher loadings of the larger particles, while in the GMA, it was the smaller range that dominated the area.

208

9.22 O3
Ozone trends for both urban areas are depicted in Figure 9.5. In this plot, the difference in concentration also becomes obvious and suggests that the processes involved in the formation of this secondary pollutant were different. One of the most important factors contributing to the difference in concentration was the level of insolation. In the MCMA, being in the tropical latitudes, the amount of sunlight hours and the intensity of the irradiation were much higher, while in the GMA both parameters were lower, also due to its latitudinal location. Figure 9.5 O3 Average in the MCMA and the GMA
O3. MCMA, 1994 - 96 O3. GMA, 1997

60

50

40

ppb

30

20

10

0 Jan Feb Mar Apr May Jun Jul Ago Sep Oct Nov Dec

The background concentration in both agglomerations were very different with the MCMA, registering a range of 35ppbv45ppbv and in the GMA the range was from 10 to 20 ppbv. Moreover, in the time series the association between NO, NO2 and O3 were also rather different. In the MCMA, photochemical formation through the photolysis of NO2 was the dominant factor controlling the concentration of O3, while in the GMA it was the titration with NO that prevailed. This can be readily seen in the respective O3 and NOx plots, Figures 5.22 for the MCMA and 7.11 for the GMA. The curves representing O3 and NO in the GMA followed a typical relationship, with high ozone loadings being related to low NO, while in the MCMA this relationship was not obvious.

209

Furthermore, in the respective discussion of the results, it was shown that the removal of ozone occurred through different mechanisms. In the GMA it was mostly through titration with NO while in the MCMA, wet chemistry was responsible for a large proportion of the removal. This can be readily seen in the respective diurnal charts, Figures 7.21, 7.22 and 7.23, 7.24 for the GMA and 5.34 and 5.36 for the MCMA. In the former plots, the variation in the NO concentration between day and night was substantial, while in the latter there was very little variation. Moreover, the relationship between NO2 and O3 during the peak times of photochemical activity was also well represented in the MCMA, with high O3 occurring when NO2 was low. As explained in the discussion of the results of this metropolis, high NO2 photolyses to form large quantities of O3, but the reaction depletes the precursor species. During the winter and autumn, the relationship was the opposite, with large quantities of NO 2 being related to large quantities of O 3 . In the GMA, the relationship between NO2 and O3 was well represented throughout the year, with low quantities of the precursor species translating into high ozone. Another important difference between ozone formation in the GMA and in the MCMA was the place of occurrence. In the GMA O3 was produced in the urban plume, in the adjacent rural areas and was transported into the urban area by the wind. Proof of this fact emanates from the diurnal charts where it can be observed that the concentrations during the night were sometimes higher than in the day, Figures 7.21 and 7.22. Furthermore, the O3 concentrations increased in a south to north trajectory, with higher concentrations in the suburban locations and lower in the city centre and other locations with high NO emission (seen in Figure 7.7). Further corroboration came from the wind analyses where it was seen that the concentration of O3 increased in the direction of the prevailing wind (Figure 7.39). In the MCMA, however, ozone was produced in the city centre and blown towards the south of the city, also consistent with the prevailing wind direction (Figure 5.46). The analyses performed therefore suggested that O3 was produced in the rural areas surrounding the GMA and in the city centre of the MCMA. Further differences were accounted in the VOC to NOx ratios. In the MCMA, it was established that the atmosphere was NOx limited, while in the GMA the relationship was determined by the availability of VOCs. The ratio in the MCMA was reported to be in the order of 13 50 (DDF, 1997) while the GMA reported a VOC limited situation (PORG, 1997). The difference is important because under NOx limited conditions, ozone concentrations are determined by the availability of NOx, a pollutant that was abundant in both urban areas. The difference in the limiting factor for the production of O3 further corroborates the dominance of NO in the tropospheric concentration of ozone.

210

As the VOC to NOx ratio determines the formation potential of O3, NOx limited conditions are tightly linked to NO2, while under VOC limited conditions, the availability of hydrocarbons determines such potential. In the MCMA ozone concentrations were determined by the photolytic reaction of NO2, while in the GMA it was the titration with NO that controlled the tropospheric concentration. When all the information is accounted for, it can be stated that the factors controlling the formation and transformation of O3 in both urban areas were substantially different. In the MCMA, an area with higher insolation and radiation intensity, the concentration of O3 was up to nine times higher than in the GMA. The relationship with NO2 dictated the tropospheric concentration in the MCMA, while titration by NO controlled the loading of O3 in the GMA. The removal mechanisms of O3 were also different, with wet chemistry playing an important role in the MCMA and titration by NO dominating in the GMA. Ozone was produced in the adjacent rural areas of the GMA, while it was formed in the city centre of the MCMA. Although O3 is only produced by the photolysis of NO2, the relative abundance of NOx and VOCs determines its formation potential. In the MCMA, the availability of NOx dictated the O3 formation potential, while in the GMA it was the abundance of VOCs. Although the process involved in the formation and removal of O3 were quite different in both urban areas, ozone can only be produced by one photochemical reaction, that of the photolysis of NO2.

9.3 Conclusion
The analyses performed in the MCMA and the GMA were compared in the preceding sections. They show that meteorological parameters affected the concentration of both PM10 and O3 in both areas. The available data suggested that the formation and transformation of both pollutants differed in several aspects, including their loadings in the two urban agglomerations. One of the most striking differences was the average loading exhibited in the two areas. The average concentration of PM10 was two times higher in the MCMA, while ozone loadings were three times higher also in the MCMA. As stated, meteorology played an important role in the concentration distribution and the loadings of both PM10 and O3. It was found that temperature, relative humidity, pluvial precipitation, and under the higher regime wind speed influenced the formation of both pollutants in the MCMA. In the GMA, however, wind speed and temperature dominated the meteorological influence. Relative humidity and rain were relatively constant and thus exhibited no relationship to concentration of either pollutant. In the MCMA PM10 were determined to be mostly primary, with only a small proportion of secondary particles.

211

The GMA, however, particles were mostly secondary and related to the emissions of motor exhausts. The chemical composition of PM2.5 showed that in both areas the role of NO2 was more important than SO2, although there was a discrepancy on the relevance of SO2 in the GMA particles. Nonetheless, it was concluded that NO 2 was the largest contributor to secondary particles in both areas. Ozone in the GMA was mainly produced in the surrounding rural areas and its concentration was controlled by its titration with NO. In the MCMA, ozone was formed in the city centre and the photolytic reaction of NO2 dictated the tropospheric concentration of O3. In both areas, ozone was transported through the cities and the highest concentrations occurred on the stations furthest away from the origin of the prevailing wind. When all the information is analysed together, several conclusions can be drawn. First and foremost that meteorology had a paramount impact on the formation and transformation of both pollutants in the two urban areas. The majority of the particles in the MCMA were primary, while they were secondary in the GMA. In both locations, the contribution of the NO2 to the formation of secondary particles was greater and SO2 played a secondary role. Carbon compounds represented a large proportion of the secondary particles in both areas. Ozone concentrations were controlled by the photolytic reaction of NO2 in the MCMA and by titration with NO in the GMA. O3 concentrations in both locations increased in the direction of the urban plume, and in the GMA most of the ozone was produced in the adjacent rural areas. Although both urban agglomerations had different factors controlling the concentration of both PM10 and O3, the relevant mechanisms including the chemistry and physics were the same. Secondary particles were produced through the oxidation of NO2 and SO2, and ozone was formed by the photolytic reaction of NO2.

212

Conclusion

Conclusion
The information presented throughout this thesis pertains to the levels of pollution in the MCMA and the GMA, especially of PM10 and O3. With this purpose in mind, the first chapter explored the influence that the different forces within the boundary layer and its structure has on the concentration of pollutants. The following chapter concentrated on the formation and transformation processes of aerosols, while in the third chapter ozone formation was explored. Throughout chapter four the two urban agglomerations were described in terms of their geographic features, their socio-demographic patterns and their history, all factors that affect the concentration of contaminants in the two urban areas. The monitoring networks of both agglomerations were described and with this in mind, the trends of PM10, O3, their precursors and meteorological parameters were presented in chapter five and seven. A thorough examination of the factors affecting the concentration of PM10 and O3 were analysed in chapter six for the MCMA and eight for the GMA. In chapter nine all the information analysed and discussed in the previous chapters was compared. The results presented in chapter five and seven were analysed through five different methods. First, a comparison of the time series of PM10, O3 and the meteorology of both areas was undertaken. The associations of both pollutants were examined through information from the time series and from the literature. Thirdly, an analysis of the diurnal trends was done to examine the influence of social activity, insolation and seasonally changes, such as temperature and pluvial precipitation. Next, simple statistical analyses were done to the data in order to quantify the relationships established. Lastly, wind direction analyses were undertaken to find out the origin and sources of both PM10 and O3 and to corroborate the associations previously determined.

Urban Areas
There are many factors affecting the pollution levels of an area, but among the most important are the geographic, socio-demographic and climatologic patterns. In both the MCMA and the GMA, these factors contribute to the distinct formation and transformation mechanisms of PM10 and O3. The basic features of both agglomerations are summarised below. The MCMA is a metropolis in the middle of a high altitude U shaped basin in the centre of Mexico. The valley is surrounded by mountains and two volcanoes, while the climatic conditions dominating the area are generally described as semi-arid. The seasonal variation of temperature is relatively small, but the diurnal range is large. The rainy season is fairly well defined, with most of the precipitation occurring during the summer. Wind speeds are relatively slow, in the order of 1 2 meters per second, and the prevailing wind direction originates from the north. All these geographic features contribute to the elevated pollution levels that the metropolitan area and its inhabitants are exposed to on a daily basis. 215

Contributing to this situation is the socio-economic structure of the country, as the MCMA houses a large proportion of the industrial base of Mexico. Most of the industrial parks are located in the north of the area, where the prevailing wind disperses most of the pollutant emissions throughout the rest of the metropolis. Furthermore, the large pool of automobiles contributes by emitting large quantities of nitric oxide, particulates and other pollutants that enhance the concentration of both PM10 and O3. These two pollutants have been described as the two most severe pollution problems in the MCMA. The Greater Manchester Area lies in a U shaped valley that opens to the west and is bounded throughout by hills that rise to over 480 metres above sea level. It is situated on the east bank of the River Irwell in south-east Lancashire and its boundaries extend beyond the River Mersey in northern Cheshire. The valley is dominated by a maritime climate that is affected by a sheltering effect from the Welsh Mountains and that results in overcasted days throughout the year. The GMA has a lower than average amount of rainfall for the region, but it is distributed throughout the whole year. Temperatures are without extremes, with a winter average value around 4 C and in the summer, the average temperature stands at about 15 C. Air pollution in the GMA has been a problem since the industrial era, and pollution control has been practice since then. It originally started as a result of its booming cotton industry, but now is more related to the transport sector. Today, the GMA houses around one million automobiles that make around six million trips per day.

Meteorology
The time series of PM10 and O3 were studied in relation to the distribution of meteorological parameters. By means of the time series analysis it was determined that in the MCMA, temperature, relative humidity, pluvial precipitation and under the higher regime, wind speed affected the concentration of both pollutants. In the case of the particulates, temperature mostly affected their loading by lowering or increasing the height of the mixed layer. This is clearly shown in the respective times series, where lower concentrations occurred during the late spring and summer and higher loadings in the autumn and winter periods. The association of temperature with the level of insolation, however, affected the loadings of ozone. In their respective series, it can be seen that the concentration of O3 and temperature climaxed during May, while they were lower during the rest of the months. Relative humidity can be directly related to pluvial precipitation, as the latter requires high humidity to occur. The loadings of both PM10 and O3 were reduced by the occurrence of rain. During the three years analysed the relationship between pollutant loadings and wind speed displayed two different modes.

216

In 1994 and part of 1995 the phenomenon of el nio affected the area causing the wind speed of the MCMA to increase to twice its norm. Under these higher speeds, the concentration of both PM10 and Ozone showed a negative relationship with wind speed, while in the rest of the period no relationship was exhibited. Throughout the time series of PM10 and O3, it became evident that temperature and wind speed dominated the meteorological influence in the GMA. Pluvial precipitation and relative humidity were relatively constant throughout the whole year and hence their impact was not readily seen in the series of either pollutant. The effect of temperature and wind speed can be readily seen in the concentration loadings of both PM10 and O3 during the first two months of 1997. During January, the temperature was relatively cold, and wind speed was relatively low. A high-pressure system dominated the area, thus the combination of low temperatures with atmospheric stability creating low wind speeds resulted in low mixed layers and the trapping of pollutants close to the ground. The stable conditions propitiated high loadings of particles and other primary pollutants including NO, thus the concentration of O3 was low (through titration by NO). During February, the temperature increased, a low-pressure system became dominant and the relatively unstable atmosphere produced high wind speeds. This in turn had the opposite effect, high mixed layers resulting in low particle loadings and high ozone concentrations, mainly for the same reasons as before.

Pollutants
As the purpose of this thesis is to examine the formation and transformation mechanisms of PM10 and O3 in the MCMA and the GMA, several factors had to be contemplated. In the case of the particles, the nature of their size distribution was considered important, as these determine their origin as primary or secondary. Their chemical composition was also important, as this would determine the relationship with their major precursors. In the case of ozone, the relationship between the NOx species was paramount as they control the amount of photochemistry and titration that occurs in the troposphere. The diurnal behaviour of ozone was very important as it showed the strength of the titrating capacity of NO. In the MCMA, through the literature, it was established that around 20% of the particles were smaller than 2.5m in diameter and were therefore categorised as secondary. The similarity of the time series suggested that NO2 was the greater contributor to this category of particles, while SO2 showed no contribution. In the chemical analyses reported in the literature, however, sulphate was attributed a relatively large proportion of the fine particulates, with only nitrate having a larger percentage. The statistical analyses corroborated the findings of this work, showing that NO2 was much better correlated to PM10 than SO2. 217

The strength of the coefficient of determination also reflected the relatively small proportion of secondary particles. Furthermore, the regressions confirmed that relative humidity had a negative relationship with the particles, therefore suggesting that rain was a major removal mechanism of PM10 in the atmosphere of the MCMA. Ozone concentrations in the MCMA were controlled by the photolysis of NO2. The respective time series of the two pollutants show that, at periods of high photochemistry, there is an inverse relationship between NO2 and O3, while a direct relationship at periods of lower irradiation. The association between NO and O3 was typical and consistent with the chemistry involved. Large quantities of NO titrated large quantities of O3, with the time series displaying the negative relationship. The diurnal analyses, however, showed that NO concentrations remained relatively constant throughout the day and the night time periods, suggesting that it was not reacting with O3 during the night. Based on this information, it was concluded that ozone was preferentially removed from the troposphere by wet chemistry. The ratio of VOC to NOx ratio reported in the literature indicated that the atmosphere is NOx limited and therefore a reduction in NOx would limit the production of O3. Examination of the wind patterns and the pollutant loadings in the MCMA, corroborated several of the hypotheses posed throughout the other analyses, that PM10 and O3 were related to their precursor species. Furthermore, they showed that the particles, NO and SO2 were largely produced in the north and were transported south. Ozone loadings throughout the city suggested that ozone was produced, to a large extent, in the city centre and transported to the south of the area by the wind. In the GMA, around 75% of the PM10 were determined to be secondary, as they were less than 2.5 m in diameter. The similarity of the time series showed that the oxidation of NO2 was the greater contributor to the particulate mass, while the oxidation of SO2 played a lesser role. In the literature, however, the chemical composition of the secondary particles showed that sulphate was attributed the larger proportion while nitrate had a secondary role. The analyses performed here, however, showed that the oxidation of NO2 was more important in secondary particle formation. In the literature, Harrison (1997) reported that in Birmingham carbon compounds represented a relatively large proportion of the particle mass. Since the two areas are relatively similar, it is stipulated that in the GMA organics formed a large proportion of the secondary particles. The statistical regressions revealed that NO2 was better correlated with the PM10 than SO2. Furthermore, the strength of the relationships reflected the relatively high proportion of secondary particles in the PM10, as the CODs were much higher than in the MCMA. The concentration of ozone in the GMA was dictated by its titration with NO and not by the photolysis of NO2. This can be clearly seen in the respective time series, where the trends of NO and O3 showed very similar patterns but in opposite directions.

218

Furthermore, when the concentration of NO was high during January, ozone was low, and when NO was low during February, ozone concentrations were high. During this same time, NO2 was also high in January and low in February, but the ozone concentration did not increase as a result of the higher NO2 loadings. Although the extreme concentrations experienced during these two months were the result of weather and meteorological conditions, it exemplifies the relationship between NO, NO2 and O3. The relationship between the three pollutants was further corroborated through the statistical analyses performed where NO was better correlated with O3 than NO2. Ozone in the GMA was produced in the urban plume and in the surrounding rural areas. Diurnal distributions showed that the O3 concentrations during the night were sometimes higher than in day time period, suggesting that it came form elsewhere. Moreover, the concentration of ozone increased in a southerly direction, and was higher at the suburban stations than in the city centre. The analyses relating wind direction to the pollutant loadings confirmed the hypotheses posed before: the association between secondary PM10 and its precursors, as the loadings originated from the same place. O3 was produced in the urban plume and its concentration increased in the direction of the prevailing wind.

Further Research
There are many areas where knowledge in the area should be enhanced. First and foremost, research should concentrate on the chemical composition of particles in both the GMA and the MCMA. These type of studies help to corroborate the species contributing to the formation of secondary aerosols. Secondly, more particle size characterisation research should done in both urban areas, as these determine many of the properties of the aerosol. Both of these types of studies are relatively scarce in the literature and are fundamental in trying to understand the formation and transformation processes of PM10. Ammonia should be monitored in the networks of both areas as this pollutant determines some of the most important gas to particle conversion processes. Loadings of this pollutant are also relatively scarce in the literature. Monitoring of methane and other reactive organic species should also be incorporated into the networks, as these determine many of the oxidation processes in the atmosphere.

219

References

References
AGMA (1997) Air Quality Management Strategy for Greater Manchester: Clearing the Air. Document Prepared by the Multi-disciplinary Officers Steering Group, May 1997. Aldape F. et al (1991) Two year Study of Elemental composition of Atmospheric Aerosols in Mexico City. International Journal of PIXE 1, No. 4:373-388. Automatic Urban Network (AUN) (1998) Concentration Data for the Grater Manchester Area. Pollutant Concentration for Seven Stations within the GMA. British Atmospheric Data Centre (BADC) (1998) Synoptic Meteorological Data. Meteorological data form Ringway (Manchester Airport) Station. Barlow M. (1995) Greater Manchester: Conurbation Complexity and Local Government Structure. Political Geography 14, No. 4:379-400. Bartholomew J. (9th Edition) The Survey Gazetteer of the British Isles, Including Summary of the 1951 Census. Published by John Bartholomew and Sons, LTD at the Geographical Institute, Edinburgh. Britannica CD-ROM (1997) Oceans: The el Nio Phenomenon. Encyclopaedia Britannica, Inc., 1997. Choularton T.W. (1999) Personal Conversation. Concentration of Ammonium in the GMA. Chow J.C. et al (1994) Temporal and Spatial Variations of PM2.5 and PM10 Aerosol in the Southern California Air Quality Study. Atmospheric Environment 28, no.12: 20612080. Cicero-Fernndez P. et al (1993) TSP, PM10 and PM10/TSP Ratios in The Mexico City Metropolitan Area - A Temporal and Spatial Approach. Journal of Exposure Analysis and Environmental Epidemiology 3, No.S1: 1-14. Clark W.A.V. and Hosking P.L. (1986) Statistical Methods for Geographers, 1st ed. John Wiley and Sons, Inc., New York. Pg. 378 380. DDF (1990) Programa Integral Contra la Contaminacin Atmosfrica. Report prepared by the Government of the Federal District, Mexico. DDF (1993) Partculas Suspendidas, Situacin Actual en la Zona Metropolitana del Valle de Mxico. Report prepared by the Metropolitan Commission to Prevent and Control Environmental Pollution in the Valley of Mexico, General Ecology Directorate. DDF (February, 1994) Informe Ejecutivo de la Calidad del Aire (1991-1993). Report prepared by the Metropolitan Commission to Prevent and Control Environmental Pollution in the Valley of Mexico, General Ecology Directorate. DDF (June, 1995) Reporte Mensual de la Red Automtica de Monitoreo Atmosfrico (RAMA). Report prepared by the Metropolitan Commission to Prevent and Control Environmental Pollution in the Valley of Mexico, General Ecology Directorate. DDF (1997) Programa para Mejorar la Calidad del Aire en el Valle de Mxico 19952000. Report produced by the Department of the Federal District, the Government of the State of Mexico and the Secretariat of the Environment, Natural Resources and Fisheries and the Secretariat of Health. DDF (1996) Informe Anual 1995. Red Automtica de Monitoreo Atmosfrico de la Ciudad de Mxico (RAMA). Derwent R.G. and Davis T.J. (1994) Modelling the Impact of NOx or Hydrocarbon Control on Photochemical Ozone in Europe. Atmospheric Environment 28, No.12: 2039-2052. Doyle G. J (1961) Self-Nucleation in the Sulfuric Acid-Water System. Journal of Chemical Physics (cited in Seinfeld, 1986) 35:795-799.

223

Finlayson-Pitts B. J and Pitts J. N (1986) Atmospheric Chemistry: Fundamentals and Experimental Techniques. John Wiley and Sons, New York, NY. Fox D.L. (1986) The Transformation of Pollutants. In: A C. Stern. (Ed.) Air Pollution, 3rd ed., Vol. VI. Academic Press, London, UK, pp. 61-94. Hare F.K. (1989) Climatology and Meteorology. In: Henry J.G. and Heinke G.W. (Eds.) Environmental Science and Engineering. Prentice Hall, Inc., Englewood Cliffs, N.J., pp. 212-250. Harrison R. M et al (1997) Sources and Processes Affecting Concentrations of PM10 and PM2.5 Particulate Matter in Birmingham (U.K.). Atmospheric Environment 31 No. 24: 4103-4117. Hinds W.C. (1982) Aerosol Technology (cited in U.S.E.P.A, 1996a). John Wiley and Sons, New York, NY. HMSO (1995) Transport Statistics for Metropolitan Areas 1995. Report Produced by the Directorate of Statistics in the Department of Transport, June 1995. de Icaza G. and Choularton T.W. (1998) Time Series Analysis of PM10 and O3 in st the Mexico City Metropolitan Area. Air and Waste Management Association, 91 th Annual Meeting and Exhibition. Conference Proceedings, June 1418 . San Diego, California, USA: CD-ROM. de Icaza G. and Choularton T.W. (1999) Particulate Matter under Ten Micrometers (PM10)and Ozone Trends in Greater Manchester During 1997. Air and Waste nd Management Association, 92 Annual Meeting and Exhibition. Conference Proceedings, th June 2024 . St. Louis, Missouri, USA: CD-ROM. IMP (Mexican Petroleum Institute) and Los Alamos National Laboratory L.A.N.L. (April 20, 1994) Proposal for a Mexico City Aerosol and Visibility Evaluation Research Project (AVER). INE (Instituto Nacional de Ecologa) (1997) Primer Informe Sobre la Calidad del Aire en Ciudades Mexicanas 1996. September 1, 1997. Jacobson M.Z. (1997) Development and Application of a New Air Pollution Modelling System-II. Aerosol Module Structure and Design. Atmospheric Environment 31, No. 2:131-144. Jacobson M.Z. (1997) Development and Application of a New Air Pollution Modelling System - Part III. Aerosol-Phase Simulations. Atmospheric Environment 31, No. 4:587-608. Jenkin M.E. et al (1996) Gas-to-particle conversion pathways. AEA Technology Report AEA/RAMP/20010010/001, Culham Laboratory, Oxfordshire, UK. Landranger (1997) Manchester Bolton and Warrington Ordnance Survey, Map No. 109, 1:50000 Scale: 1:50000. Logan J.A. et al (1981) Tropospheric Chemistry: a global perspective. Journal of Geophysic Research (cited in Seinfeld and Pandis, 1998) 86:7210-7254. Lyons T.J. and Scott W.D. (1990) Fundamentals of Air Pollution. Belhaven Press, London Mann C. (1994) Changing Air Quality in Greater Manchester. Atmospheric Research and Information Centre ARIC Pamphlet: 1-8. MCC (Manchester City Council) (1997) Air Quality Monitoring Report 1996/97. Report undertaken by the Pollution Group of the Environmental Health Division of the City of Manchester. Report Ref. JDD/RHL/AP97. McRae G.J. et al (1982) Development of a Second-Generation Mathematical Model for Urban Air Pollution I. Model Formulation. Atmospheric Environment 16:679-696.

224

Miranda J. et al (1994) Determination of Elemental Concentrations in Atmospheric Aerosols in Mexico City Using Proton Induced X-Ray Emission, Proton Elastic Scattering, and Laser Absorption. Atmospheric Environment 28 No. 14:2299-2306. Miranda J. et al (1992) A Study of Elemental Contents in Atmospheric Aerosols in Mexico City. Atmsfera 5:95-108. Oke T.R. (1987) Air Pollution in the Boundary Layer. In: Boundary Layer Climates, 2nd ed. Routledge, London, UK, pp. 304-338. Oke T.R. and Jauregui E. (1992) The Surface Energy Balance in Mexico City. Atmospheric Environment 26B, No. 4:433-444. Pandis S. N and Seinfeld J. H (1989) Mathematical Modelling of Acid Deposition Due to Radiation Fog. Journal of Geophysics (cited in U.S.E.P.A, 1996a) 94:12:911-923. PORG (1997) Ozone in the United Kingdom. Forth Report of the DETR Photochemical Oxidants Review Group, 1997. Department of the Environment Transport and Regions. RAMA (Red Automtica de Monitoreo Atmosfrico) (1997) Data for the Mexico City Metropolitan Area. Concentration Data and Meteorological Parameters for Five Stations within the MCMA. Ravetz J. (1997a) City-Region 2020. Integrated Planning for Long-term Sustainable Development in a Northern Conurbation. Part B Chptr. 3 The State of the City-Region. 3rd edition of Consultation Draft. Ravetz J. (1997b) City-Region 2020. Integrated Planning for Long-term Sustainable Development in a Northern Conurbation. Part D Chptr. 8 Travel and Transport. 3rd Edition of Consultation Draft. Ravetz J. (1997c) City-Region 2020. Integrated Planning for Long-term Sustainable Development in a Northern Conurbation. Part D Chptr. 10 Waste and Pollution. 3rd Edition of Consultation Draft. RMS Royal Meteorological Society U.K. (1997a) Weather Log for January, 1997. Weather 52 No.3. RMS Royal Meteorological Society U.K. (1997b) Weather Log for February, 1997. Weather 52 No.4. RMS Royal Meteorological Society U.K. (1997c) Weather Log for June, 1997. Weather 52, No.8. RMS Royal Meteorological Society U.K. (1997d) Weather Log for September, 1997. Weather 52 No. 11. Secretara de Salud (December 3rd, 1994) Air Quality Standards for Air Pollutants. Published in the Official Diary of the Federation. Seinfeld J.H. (1986) Atmospheric Chemistry and Physics of Air Pollution. John Wiley and Sons, Inc., New York, NY. Seinfeld J.H. and Pandis S.N. (1998) Atmospheric Chemistry and Physics: From Air Pollution to Climate Change. John Wiley & Sons, Inc., New York, NY. Streit G.E. and Guzmn F. (1996) Mexico City Air Quality: Progress of an International Collaborative Project to Define Air Quality Management Options. Atmospheric Environment 30 No. 5:723-733. Stull R.B. (1988) An Introduction to Boundary Layer Meteorology. Kluwer Academic Publishers, Dordrecht, The Netherlands. U.S. EPA (1996a) Air Quality Criteria for Particulate Matter, Vol. I. National Center for Environmental Assessment, Office of Research and Development (EPA/600/P95/001aF), North Carolina.

225

U.S. E.P.A. (1996b) Air Quality Criteria for Particulate Matter, Vol. III. National Center for Environmental Assessment, Office of Research and Development (EPA/600/P95/001cF), North Carolina. Vega E. et al (1997) The Use of a Receptor Model for Fine Particulate in Mexico City. Air and Waste Management Association, 90th Annual Meeting and Exhibition. th Conference Proceedings June 8-13 . Toronto, Canada: CD-ROM. Villalobos-Pietrini R. et al (1995) Mutagenicity Assessment of Airborne Particles in Mexico City. Atmospheric Environment 29, No. 4:517-524. Warneck P. (1988) The Atmospheric Aerosol. In: Chemistry of the Natural Environment. Academic Press, London, UK, pp. 278-373. Wayne R.P. (1991) Chemistry of Atmospheres, 2nd ed. Oxford University Press, Oxford. Weinstock B. et al (1980) Chemical Factors Affecting the Hydroxyl Radical Concentration in the Troposphere. In: Pitts J.N. and Metcalf R.L. (Eds.) Advances in Environmental Science and Technology. John Wiley & Sons, Inc., New York, NY, pp. 221-258. Willeke K. and Whitby K. T (1975) Atmospheric Aerosols: Size Distribution Interpretation. Journal of the Air Pollution Control Association (cited in U.S.E.P.A, 1996a) 25:529-534. Williams G. (1996) City Profile, Manchester. Cities 13 No. 3:203-212. Williams M.D. et al (1995) Development and Testing of Meteorology and Air Dispersion Models for Mexico City. Atmospheric Environment 29, No. 21:2929-2960. Wood C.M. et al (1974) The Geography of Pollution. A Study of Greater Manchester. Manchester University Press, Manchester. Yocom J.E. et al (1986) Air Pollution Effects on Physical and Economic Systems. In: Stern A.C. (Ed.) Air Pollution, 3rd ed., Vol. VI. Academic Press, London, UK, pp. 145-157.

226

Formation and Transformation Mechanisms of Particulate Matter Under Ten Micromilimeters (PM10) and Ozone (O3) in the Mexico City Metropolitan Area and the Greater Manchester Area se termin de imprimir en Febrero de 2003. Composicin tipogrfica, edicin e impresin Comercializada Collage Creativo S.A. de C.V. Agustn Lara Mz. 8 Lote 18, Col. Hogar y Redencin C.P. 01450, Alvaro Obregn, Mxico, D.F. Tel: 56-43-87-81 Se tiraron 350 ejemplares ms sobrantes para reposicin.

Você também pode gostar