PHYSICS AND CHEMISTRY OF ATMOSPHERIC AEROSOLS A COURSE GIVEN AT THE DEPARTMENT OF METEOROLOGY, UNIVERSITY OF STOCKHOLM, FALL SEMESTER, 1976. BY RUDOLF B HUSAR 1 1. AEROSOL PHYSICAL PROPERTIES Aerosols. The term aerosol refers to liquid or solid particles suspended in the air. Aerosol particles may be liquid droplets, aggregates of odd shape, or single crystals of regular, say cubical, or irregular shape. Their chemical composition may also vary from dilute water solution of acids or salts, organic liquids, to earth's crust materials (sand) and toxic heavy metals (Pb, Hg, As). Depending on their origin and visual appearance aerosols have acquired different names in the everyday language. Dust generally refers to solid airborne material, dispersed into aerosol from grainy powders. Fumes are produced by various industrial processes such as in a foundary, welding etc.. Combustion processes produce smoke particles, but the incombustible residue of coal is also called flyash. Mist is formed when a vapor condenses to form fine liquid droplets, which may then grow to larger fog droplets. In the early days, air pollution had the appearance of both smoke and fog, so it was appealing to create a new word for it: smog. In the open atmosphere, the visibility may often be reduced by haze, originating from natural or anthropogenic sources. It is obvious that in the context of physical and chemical treatment of atmospheric aerosol behavior, it is advantageous to abandon the above set of rather poorly defined terms. Instead, the generic term aerosol is used to express all types of airborne matter in the 10 to 100 m size range. If further classification is needed, then it is beneficial to perform the subdivision based on physical or chemical grounds. In case of atmospheric aerosols it is useful to divide the aerosol population into fine particles, having a diameter less than about 2 m and course particles with size 2 m. The two aerosol populations have generally distinctly different physical characteristics (shape, volatility) and chemical composition. Furthermore, the two populations have different sources and more important - different effects. A scientific rationale for the separate consideration of fine and coarse particles is given later. 2 Table 1. Characteristics of particles as a function of their size. 3 Particle size. The size of atmospheric aerosols ranges over five orders of magnitude. The smallest observed particles are of the order of 10 (1 = 10-8 cm) in diameter and they could well be called molecular clusters. In Table 1, the typical size of diatomic molecules is indicated to be between 3 and 6 . On the other extreme, large dust particles of 100 m (1 m = 10-4 cm) may be kept suspended in the air for extended period of time. Within the five decade ranges, there are "windows" in which the particles are comparable in size to the mean free path of atmospheric gases (0.06 m), wavelength of the visible solar radiation (0.4-0.1 m), the wavelength of the outgoing infrared radiation from the earth (6-40 m) (8-14 m - atmospheric window- spechrae region over which thermal emission from ground is radiated directly to space. Paetridge - Plalt, 8, Fig. 7.1) and other important micro-physical length scales. This is of importance, because the interaction of the particle with the surrounding gaseous medium depends primarily on the ratio of the mean free path, 1, to the particle diameter, Dp, called the Knutsen number Kn. = 1/Dp. For Dp <<1 the gaseous molecules that are bombarding the particle arriving from far distance and each collision causes a substantial change in the direction or speed of the particle. Accordingly, such a particle behaves as if it were of giant molecules and its motion (drag force) and Brownian diffusion may be described by the rigorous kinetic theory of gases. Such an aerosol is said to be in the free molecular regime. On the other hand, for a particle which is large compared to 1, i.e. Kn <<1, a change in its momentum and only the combined effect of many particles yields an effective change. From the point of view of the particle, it "feels" as if it were suspended in a continuum and for this reason such a particle is said to be in the continuum regime. The interaction of a particle with electromagnetic, including visible, radiation depends largely on the ratio of the particles diameter to the wavelength, , of the incoming radiation. This dimensionless ratio, the optical size parameter, = Dp/, determines whether the interaction is purely dipole (<<1) or it is governed by geometrical optics i.e. diffraction, refraction and reflection (>>1). In addition to the particle size, the behaviour of atmospheric aerosols also depends on their shape and chemical composition. Particle shape. Atmospheric aerosols exhibit a variety of shapes. Most commonly, they are spherical, liquid droplets. The water solution of sulfuric aced or other salts that constitutes a large fraction of the fine particle mass (e.g. Brosset, 1976) is believed to be in droplet form. Photochemically produced aerosol, such as the Los Angeles smog aerosol consists of droplets formed around a nucleus (e.g. Husar et al. 1976). Several of the common aerosol salts ((NH4)2SO4, NaCl) exhibit deliquescence (June, 1963; Covert et al. 1974). This behaviour is manifested by a rapid transition from crystalline form at low relative humidity to a droplet at high RH. Dry sodium chloride crystals are almost cubical; ammonium and potassium sulfate crystals are also generally cubical but with more rounded edges. 4 Combustion and recondensation of weakly volatile substances lead to the formation of chain aggregates of solid spherical spheres. The aggregation is generally taking place via coagulation in the vicinity of the combustion zone and the individual spheres in the chain are held together by molecular forces. When suspended in the air, a chain aggregate may bend, twist and rotate due to asymmetric molecular bombardment of fluid shear. A sample of atmospheric fine particles impacted onto a grid and viewed through electron microscope is shown in Figure 2. The chain aggregate in the center, probably originated from an automobile exhaust. Unfortunately, atmospheric liquid particles lose their identity when deposited onto a grid, inserted into the vacuum of the electron microscope and exposed to the heating due to the absorbed electrons. However, such particles leave a flat, circular residue as evidence of their original form. Size segregated samples of coarse atmospheric particles are shown on the scanning electron micrograph (Figure 3). They are generally non-spherical with major fraction resembling soil dust or fibrous matter. The characteristic size of spherical particles is their diameter. Both the light scattering and mechanical (e.g. fluid resistance) properties of nonshperical particles depends strongly on their particular shape. However, the definition of a characteristic size is difficult. The method of their measurement is often used as the definition of their size. For example, in case of a cascade impactor, particles are sorted according to their aerodynamic (or Stockes) diameter which encompasses all the particles having the same trajectory in a curved flow field. A small but dense particle may have the same aerodynamic size as a large particle with low density or an aggregate with large aerodynamic resistance. Sorting or sizing of particles with optical single particle counter/spectrometers, yields an "equivalent optical diameter". Here again, a multiplicity of physical sizes may correspond to the same equivalent optical diameter if their refractive indexes are different. The above discussion points to the need to specify the method of measurement whenever particle size or size distribution are presented. It is sometimes useful to express the shape of an aerosol by shape factors. For a particle with given geometrical shape, the shape factors for volume, v, and shape factor for surface, s, are defined as: V = v Dp3 s = s Dp2 For volume, v and surface, s, respectively. Dp refers to a typical length scale of a particle. The shape factors of several common shapes are given in Table 1. 5 Figure 1. Transmission electron micrograph of size classified submicron aerosols. The white area in the center is where the collodion film is broken and the chain agglomerate aerosol is directly exposed to electron beam. Most of the submicron aerosol volume is made up from liquid droplets (which leave circular residues) and of chain agglomerates. Figure 2. Scanning electron micrograph of size classified (greater than 2 m) atmospheric aerosol collected by an impactor. The collection substrate is a collodion coated electron microscope grid. Note the irregular shap of the particles. 6 Table 1. Shape factors for surface and volume. v s Sphere /6 = 0.52 Cube 1 6 Crushed sand 0.28 2 Calcite 0.32 2.5 Particle concentration. The aerosol concentration is defined in a similar manner as the density: the number N or mass M of particles per unit volume, V, of air (N/V). the concentration will vary, depending on the size of the volume element V. At large V, the concentration, say the total number of particles N, will vary from one location to another because of macroscopic gradients. With decreasing size of V the number of particles may decrease such that it may exhibit statistical fluctuations because of the finite number of particles. Mass concentrations of atmospheric particulate matter may reach several hundred micrograms per cubit meter. This corresponds to a mass ratio of the order of 10-7 grams of particulate matter per gram of air. Thus the contribution of the aerosol to the total mass, even in highly polluted atmospheres, is negligible. For most applications, the effect of the particles on the fluid motion can be neglected. The total number concentration may range from 102 particles/cm3 in a very clean atmospheric air to 106 particles/m3 near intensive combustion aerosol sources such as roadways. Even at 106 cm-3 the average distance between individual particles is about 100 m, which is several orders of magnitudes larger than the characteristic size of combustion aerosols (0.01-0.1 m). Therefore the motion of aerosols is free from mutual interaction, except when they approach each other and adhere due to molecular forces (coagulate). Size distribution. The most important physical characteristic of an aerosol population is the particle size distribution. In principle, the aerosol distribution function should have two independent aerosol variables, viz. particle size and chemical composition, as well as the usual space and time co-ordinates, x, y, z and t. in practice, it is convenient and adequate to treat the chemical distribution as discrete set of species, rather tan as a continuos variable. Thus we may define a distribution function ni; such that ni (r, x, y, z, t) represents the number concentration of chemical species I, at position x, y, z and time t, and in the range of particle radii between r an r + dr. At times it is instructive to consider higher weightings of the number distribution function such as the distribution of aerosol surface and volume (or mass) with respect to size. The surface, volume and mass distribution functions, si, vi and mi respectively, are related to the number distribution as follows: si = 4 r2 ni weighting function 7 vi = 4/3 r3 ni mi = 4/3 i r3 ni where i is the mass density of species i. The total aerosol distribution function in each case may be expressed as the summation over all chemical species. Thus, for example, the total mass distribution function is defined as m(r, x, y, z, t) m i (x, y, z, t, r) i Typical total distribution functions for number n, surface s, and volume v, for the Los Angeles smog is given in Figure 4. Figure 3. Average size distribution of the Los Angeles smog aerosol (Whitby, Husar & Liu, 1972). Integral Moments of the Distribution Function. The effects of a polydisperse aerosol population are best described in terms of integral moments of the size spectrum i.e. integrals of the size spectrum weighted by a function of particle size. If the weighting function is an integer power law (many people confuse power law (xa) with exponential functions e-x) function of r, then the v-the integral moment, Iv, is defined as I v r ν n(r)dr 0 8 Integral moments have physical significance describing the behaviour of a polydisperse system as described by the following examples. The zeroth moment is the total number concentration, N, defined as I 0 n(r)dr N 0 The first moment I1 rn(r)dr 0 gives the total length of a chain if particles were lined up next to each other. A physically more meaningful parameter is the size (radius) of the average particle, r, defined as I1 Io rn(r)dr 0 r n(r)dr 0 The second moment is proportional to the total aerosol surface area per unit volume, 4I 2 4π r 2 n(r)dr S 0 S = 4I2 The surface area of the average particle is given by 4I2/Io = S/N. The third moment is proportional to the total volume, V, of the aerosol suspension per volume of air is, 4 3 πI 3 4 3 π r 3 n(r)dr V 0 The total mass, M, is M 4 π ρ(r)r 3 n(r)dr 3 0 9 where ( r ) is the size dependent particle density. For = const., M = V. The volume of an average particle, V 4 πI 3 3 VN I0 The fifth moment is proportional to the mass flux, or deposition rate, D, of material sedimenting from air (g/cm2, sec) 8π 2 gI 5 4 D v s (r)(ρ πr 3 )n(r)dr 3 27μ 0 where the sedimentation velocity, vs = 2r2g/9, and is the dynamic viscosity of the air. There are several other important moments of the size spectrum for which the weighting function can not be conveniently expressed as integer power-law of particle size. In that case the integral moment is defined as I f(r)n(r)dr 0 Notable examples for such weighting functions include the light scattering crossection, which yields the total light scattering of a polydispersed system. The relative contribution of the different part of the size spectrum to a given moment is determined by the weighting functional. In a polluted urban atmosphere the number concentration or zeroth moment is dominated by particles in the 0.01-0.1 m size range; the volume or mass concentration is contributed from both the 0.1-1.0 m and 0-30 m size range as shown in Fig. 4. Relationship Between Different Forms of Size Distribution. The numbersize spectrum of atmospheric particles extends over five orders of magnitude in particle size and up to ten orders of magnitude in concentration (Junge, 1963; Clark & Whitby, 1967). This extremely wide range coupled with the different roles of small and large particles lead to the use of several forms of distribution functions. In this section, attempt is made to facilitate a convenient intercomparison of several commonly used distribution functions. The extent of the aerosol size spectrum over several orders of magnitude necessitates the use of a logarithmic scale for particle size presentation. It is also advisable, that the area under a distribution function, when displayed graphically, is proportional to the integral (i.e. a moment) of that function. (Junge, 1963; Berry, 1967; Whitby et al., 1972)/ 10 If that criterion is satisfied, then simple visual impection of a graph reveals the value of that integral (total number, mass) as well as the contributions of different size ranges to the integral. Such presentation is facilitated by the semi-log plot of a properly defined spectrum function (linear) vs. log of particle size (Berry, 1967; Whitby, 1972). If the distributed variable is. For example the total number such distribution function, n(log r), is defined as, dN = n(log r) d(log r) where dN is the number of particles in the radius range between log r and log r + d (log r). Conventionally decimal logarithms are used. Unfortunately, practical problems impose another problem of presentation. The measurements of aerosol size or size distribution is most frequently performed in terms of particle diameter, Dp, rather than radius. The corresponding number distribution function is n(log Dp). In this case the numerical values of the two distribution functions are identical as shown below. Following the definition of distribution functions, that the number of particles, dN, in an infinitesimal range of the independent variable, dr, and d(log r) must be identical. So we may write dN = n(r) dr = n(log r) d(log r) Thus the transformation from one independent variable, x, to another, z, follows the equation y(z) = y(x) J where the Jacobian transformation function J = dx/dy. The Jacobian of the transformation from r to log r is J = dr/d(log r) = r/log e. Thus n(r)dr is related to the distribution function n(log r) as follows: n(log r) r n(r)dr log e 11 Table 2. Following similar transformation, the functional relationship between the most commonly used distribution functions are given in Table 2. The functions in the table are the multipliers of f2 to yield f1. For example a transformation of the number radius distribution n(r) to the volume distribution with (log Dp) is: 4r 4 n(r) 3log e For practical reasons, in the following text, we shall almost exclusively be using (log Dp) as the size variable. v(log Dp) Commonly Used Distribution Functions. There is no universal distribution function that adequately describes the majority of aerosol populations, including their various weightings. It is often desirable, however, to approximate the form of an aerosol spectrum with an analytical function. In the past , the log-normal distribution, the power-law (Junge) distribution, the gamma function and the "bimodal" distribution have been used most frequently. A. Log-normal distribution. 12 Crushing and atomization processes often give skewed distributions over broad size range as shown schematically in Figure 4. Figure 4. Features of the log-normal distribution. When such distribution are plotted on semi-log paper their shape resembles that of the normal Gaussian distribution. Functionally, the only difference from the Gaussian distribution is that the independent variable, r, is replaced by (lnr) or (log r). Thus the log-normal distribution is defined as n log r d log r (log r log rg ) 2 1 exp d log r 2(log g ) 2 2 log g where rg is the geometrical mean size and g is the geometrical mean deviation. In analogy to the calculation of the mean radius (p. 12) the geometrical mean radius rg is obtained as follows: log rg log r n log r d log r 0 Similarly, the logarithmic standard deviation is defined as log g log r log rg 2 nlog r dlog r 2 0 13 The mode of a distribution is at size r, where the maximum of the distribution occurs. The median of the distribution is defined as the size which divides the total population into two equal halves (see Figure 5 a). by definition, the median radius, rm is such that rm 0 rm f log r d log r f log r d log r where f(log r) may be any number surface or other distribution function. The mass median radius has often been used to indicate the characteristic size at which the bulk of the aerosol mass is contained. The median is best determined from normalized cumulative plots of the quantity (number, surface, volume or mass) vs. log r. Such plots are also referred to as logprobability graphs. Figure 5. Normalized cumulative distributions. Log-probability plot. A convenient property of this function is that of n(r) is log-normal, then all the other weighting, rn( r ) are also log-normal with the same g. In fact, it can be shown that the geometric mean radius log rg of a higher (-th) weighting, can be calculated from the geometric mean size of the zeroth weighting log rg0 log rg = (log g)2 + log rg0 As an example, if geometric number mean radius is rg = 0.1m and g = 2.0 then the volume mean diameter is log rg,3 = 3(log 2)2 + log(0.1) = -0.728 rg,3 = 0.187 m It is worth pointing out that the integral of the log-normal distributions unity. Such a distribution is defined by three parameters, the mean size, rg, standard deviation and a scaling factor for the concentration. Power-law distribution. In the early stages of atmospheric aerosol science, (Junge, 1963) atmospheric aerosol spectra have been fitted with a power-law function 14 n(r) = ar- a = constant where the exponent ranges from 3 < < 5. Such fits were arising as straight lines in plots of n(log r) on log-log paper. The power-law fit brought to the researcher's attention some of the regularities exhibited by the size spectrum of atmospheric aerosols (Friedlander, 1951; Clark & Whitby, 1961). More detailed examination of aerosol size spectrum (Whitby et al., 1972), in particular of its higher weighting (surface ad volume), revealed that simple functions such as the power-law, are inadequate to describe the complex and dynamic behaviors of atmospheric aerosols. For this reason, the use of power-law distributions is no recommended. A direct comparison of different distribution functions is given later. The modified gamma function. The calculation of radiation transfer through have layers, on a prior assumption needs to be made about the aerosol size spectrum. For such radiation calculations, a modified gamma function has been proposed by Diermendian (1969) and used extensively (e.g. Dave, 1971). The form of the function is n(r) = a1ra2 exp(-b1rb2) where a1, a2, b1,b2 are positive constants. a1 a2 b1 b2 Haze M 5.33104 1 8.94 1/2 Haze L 4.97106 2 15.11 1/2 Cumulus C 2.38 6 3/2 1 1.0810-2 8 1/24 3 Corona cloud C B. Bimodal distribution. Measurements of atmospheric aerosol size distributions in urban areas indicate a frequent occurrence of a bimodal distribution for the aerosol volume (mass) concentration. The two modes of the volume distribution are illustrated in Figure 4. There is accumulating evidence that the aerosol populations in the two mass modes are formed by different mechanisms, come from different sources and have drastically different chemical compositions. The two modes are separated by approximately one decade of particle diameter or more, while the width of the two distributions generally does not exceed g = 2.5. This allows a separation such that the overlap of the two modes is less than 10-20%. Since the two modes of aerosol mass are contributed by different sources, they are practically decoupled from each other in their spatial and temporal pattern. The existence of bimodal aerosol volume distributions was first pointed out by Whitby et al, (1972), while some of the physical mechanisms that shaped the bimodal distribution were discussed by Husar et al, (1972). They have compiled a number of their own size spectrum measurements obtained in Los Angeles by electrical mobility analyzer and an 15 optical counter and compared whit data of other investigators obtained in variety of locations (Figure 9). The data shown in Figure 9 indicate the existence of two aerosol populations, the fine and the coarse particles. It should be noted, however, that the relative magnitudes of the two modes (the integral under the two modes) may vary drastically from location to location. At present there is no generally accepted expression (formula) which would adequately describe the bimodal distribution. However, visual inspection of volume or mass distributions such as shown in Figure 4 and 9 suggests that the two modes of the volume distribution can be reasonably well approximated by two log-normal distributions. We have pointed out previously that the size and the size distribution of a polydispersed aerosol population depends on the method of measurement. It is, therefore, important to test whether the bimodal mass distribution emerges when the size distribution is measured by other methods. Recently, Kadowaki, (1976), reported mass distribution data obtained by a cascade impactor. His data are shown in Table 10 for comparison. The normalized cumulative form of the mass distribution function is shown in Table 11. The mass measurements of the individual stages of the cascade impactor, the size ranges for each stage, and the calculated mass distribution function are given in Table 3. The above data clearly indicate the existence of the bimodal distribution when the aerosol is characterized in terms of its aerodynamic size. Figure 6. Comparison of volume distributions measured by several investigators in different locations: Additional data are given in Table VIII. Note the universal bimodal nature of all of these data and that the data obtained by Clark, Peterson, 16 and from the more recent Los Angeles and Colorado studies, were obtained under pollution-free conditions such that it may be assumed that a background aerosol was being measured, is rising sharply at 10 m. (Whitby, Husar, & Liu, 1972). Figure 7. Histogram and size distribution curve of total aerosols in Nagoya (16-22 May 1974. Conc 131 gm3). Figure 8. Typical cumulative particle size distribution of total aerosols in Nagoya. (Kadowaki, 1976). 17 Stage Weight of total Concentration logp of total aerosols aerosols M (mg) Dp m (g m3) m logp Dp(g m) 3 0 2.18 12.2 1.00 12.2 1 1.81 10.2 0.451 22.5 2 2.51 14.1 0.398 35.4 3 2.62 14.7 0.350 42.0 4 1.86 10.5 0.451 23.2 5 1.70 9.55 0.646 14.8 6 2.27 12.8 0.525 24.3 7 2.76 15.5 0.412 37.6 Backup filter 5.59 31.4 1.68 18.7 Total 23.30 131 Table 3. Determination of histogram size distribution curve of total aerosols. Sampling period: 16-22 May, 1974; Air volume sampled: 178 m3 (Kadowaki, 1976). C. Comparison of different common distributions. The primary criterion for evaluation the quality of the fit by a model distribution function and real data is that the fit is good in the size range where a give effect (visibility, health, soiling, etc.) is of importance. For example, for light scattering calculations in the visible range the model function needs to be best in the 0.1 - 1.0 m range since that subrange contributes most to the light scattering. On the other hand for the calculation of soiling effects (deposition flux) the model distribution and the actual data have to compare well in the 5 - 30 m range. It is, therefore not likely that a single distribution function of a simple form (log-normal, power law, gamma function) will satisfy the above set criteria over the size range of five decades. It is instructive to compare graphically the shape of the above discussed distribution functions when plotted in different coordinate systems. In Table 12 we show the lognormal, power law ( = 4), and modified gamma function (haze L as defined by Diermendian). For comparison the data for the number distribution function of the Los Angeles smog aerosol are also shown. It is evident that in such a log-log plot the power law distribution is a fair approximation over a more than a decade of particle size, while the other two distributions deviate strongly both at the upper and the lower end of the spectrum. The same distribution functions and aerosol data are shown on a semilog plot of v (log Dp) in Table 13. As seen there the power law distribution with = 4 is a straight line with v(log Dp) = const. It is evident that the power law is an inadequate fit for the volume distribution in the 0.1 - 1.0 m range. A log-normal distribution with g = 2.24 and Dp = 0.3 um is a reasonably good fit for the finer particle volume mode. The shape 18 of the modified gamma function is quite similar to the log-normal distribution. However, with the constants given by Diermendian the mean size of the widely used haze L distribution is by factor of three higher than that of the Los Angeles smog aerosol. Figure 9. Grand average number size distributions for Los Angeles and the 1966 Clark study in Minneapolis: Also shown are a few data obtained by peterson in Minneapolis in 1967 under inversion conditions. The Colorado data was obtainde in Ft. Collins during the summer of 1979 under conditions where it is believed to represent pollution-free contintetal background. For comparison, the Junge distribution fitted to Clark's 1966 Minneapolis data is also shown. (Whitby, Husar, & Liu, 1972). 19 Figure 10. Grand average volume distribution from Los Angeles. The grand average volumes for subranges V3- (all particles less than 1.05 m) and for subranges V4+ (all particles greater than 1.05 m) are also shown. Also shown ist the best-fit log normal distribution to subranges V3-. The geometric mean size and geomtric standard deviation shown are for the fitted log normal distribution. (Whitby, Husar, & Liu, 1972). In conclusion we should note that for calculations involving model aerosol size distributions one should avoid using "generally accepted" distributions functions. Rather, it is best to examine the available experimental data for the particular size range of interest and arrive at a best fit function in that manner. 20 REFERENCES W. Clark & K.T. Whitby. J. Atm. Sci. (1967). D.S. Covert, R.J. Charlson & N.C. Ahlquist, "A Study of the Relationship of Chemical Composition and Humidity to Light Scattering by Aerosols". J. Appl. Meteor. 11, 968 (1972). J.V. Dave & P. Holpern, "Effect of changes in ozone amount of UN radiation received at sea level of a model atmosphere." Atm.Env. 10 547-555 (1976). D. Diermendian, "Electromagnetic scattering on spherical polydispersions". American Elsevier, N.Y. (1969). S.K. Friedlander, "On the particle size spectrum of atmospheric aerosols". J. of Meteor. 17, 373-374; 479-483 (1960). S.K. Friedlander, "Chemical Element Balances and Identification of Air Pollution Sources". Env. Sci. Technol. 7, 235-240 (1973). G. Gartrell & S.K. Friedlander, " Relating Particulate Pollution to Sources: The 1972 California Aerosol Characterization Study". Atm. Env. 9, 279-299 (1975). R.B. Husar, K.T. Whitby & B.Y.H. Liu, "Physical mechanisms Governing the Dynamics of the Los Angeles Smog Aerosol". J. Coll. Interface Sci. 39, 211-224 (1972). R.B. Husar, W.H. White & D. Blumenthal, "Direct evidence of heterogeneous aerosol formation in the Los Angeles smog ". Evn. Sci. &Technology 10, 490-491 (1976). C.E. Junge, "Air chemistry and Radioactivity", Academic Press (1963). Kadowaki, Atm. Env. (1976). K.T. Whitby, R.B. Husar & B.Y.H. Liu, "The Aerosol Distribution of the Los Angeles Smog", J. Coll. Interface Sc. 39, 177-204 (1972). 21 HOMEWORK PROBLEMS OCT. 8, 1976 PHYSICS AND CHEMISTRY OF ATMOSPHERIC AEROSOLS (Homework to be returned Mon. Oct. 18 1. Complete the conversion of distribution functions in the table on page 23 of class notes 2. Take the mass distribution function of Kadowaki (1976) (Table 3, Fig 10,11). a. Plot the mass distribution on log-probability paper (supplied) and fit the fine particle mode and the coarse particle mode each with a log-normal distribution. b. Estimate the mass median diameter and g for each mode from cumulative distribution plots. c. Assume density of 2 g/cm3 for coarse particle mode, 1 g/cm3 for fine particle mode. Calculate and plot (semilog paper) the distribution of number, surface volume and deposition flux for each size range. d. Calculate the total number N, cm-3; Surface area, S, m2/cm3; volume, V, m3/cm3; and deposition flux, D, g/cm2, sec. 22 AEROSOL MECHANICS 1. Introduction 2. Fluid Mechanical Regimes 3. Drag Force 4. Settling 5. Impaction 6. Brownian Motion 23 1. AEROSOL MECHANICS Introduction The mechanical interaction of particles with gas molecules, bounding surfaces, and with each other is described by aerosol mechanics. Particles suspended in air are constantly subjected to bombardment by N2, O2, H2O and other air molecules. The forces associated with this molecular bombardment are in general in excess of the gravitational force, hence fine particles remain suspended, such that their settling velocity is small compared to their velocity caused by other forces. When a particles is in motion relative to the surrounding gas molecules there is an asymmetry of the molecular bombardment intensity: molecules impinging near the leading edge of the particle exchange more momentum (due to higher relative impact velocities) than those impinging at the trailing edge of the particle. The overall effect is a drag force directed opposite to the direction of motion. A major component of aerosol mechanics is concerned with the determination of the drag force as a function of particle size. Motion relative to the surrounding gas molecules may be caused by a variety of forces; gravitational, inertial, electrical, and "phoretic" forces caused by gradiance of temperature, vapor pressure, etc. (Fuchs, 1964, Davies, 19 ). Two major forces acting on atmospheric aerosols are gravitational and inertial, the former being responsible for aerosol removal by settling and the latter for the inertial deposition by impaction onto ground surfaces, cloud and rain droplets. A peculiar feature of aerosol behavior is that the particles have a relative motion to the gas molecules even in the absence of external forces. Such, Brownian motion is caused by the asymmetry of the molecular bombardment at any instant of time. For instance in a short time interval the number of molecules impinging on the particle on the left side is greater than from the right side than the particle will tend to move toward the right. The net motion caused by such movement is random in stochastic sense and the entire process of random walk is referred to as Brownian diffusion. Brownian diffusion is in many respects analogous to molecular diffusion and leads to collisions between the particle and other surfaces or to collisions among particles themselves. At this point, however, the analogy between gases and particles ceases. When a small particle approaches another solid or liquid surface within a distance of several molecular sizes the particle and the surface will adhere due to the strong van der Waol's forces. For example, the strength of this attraction may easily hold a 30 m particle against the gravity. For fine particles this force exceeds gravity by several orders of magnitude, making the separation of a deposited fine particle almost impossible. Collision of particles among themselves and the subsequent adhesion is called coagulation. In case of coagulation solid particles collision leads to an aggregate, while collision of droplets leads to their merging into a single droplet or coalescence. Another area of aerosol mechanics is resuspension. This process refers to the mechanical separation of particles from the earth surface by high air velocities (shear near the surface. Atomization is the mechanism by which bulk liquid meter, e.g. 24 seawater is broken up into spray droplets. In the following sections the above processes and mechanism will be discussed in more detail. Fluid Mechanical Regimes The fluid mechanics of aerosols describes the momentum exchange between a particle and suspending gas, the overall force that arises due to molecular impingements, as well as the particle velocities and trajectories as a result of the combined interaction of all forces. The qualitative description of the gas particle interaction may be aided by the sue of a length scale characteristic for the gaseous molecules: their mean free path, 1. The mean free path is the average distance traveled by a molecule before collision with another "air molecule" (We should note here that "air molecules" refer to proper mixture of N2, O2, H2O and other molecules, each having its own mean free path depending o n molecular mass.). For air at normal temperature and pressure 1 = 0.066 m and it is inversely proportional to pressure (Fuchs, 1964). The gas-particle interaction may be divided into several regimes depending whether the particle size is much smaller, much greater or comparable to the mean free path of air. The dimensionless parameter which is used to define these regimes is the Knudsen number Kn = 1/Dp = mean free path/particle diameter The flow regimes as defined by the Knudsen number are listed in Table 1, which also contains expression for the drag force for each of the regimes. Table 1. Flow regimes defined by the Knudsen number. Regime Kn range Drag Force Dp range, m Continuum 0 - 0.1 0.6 6 rv Transition 0.1 - 7 0.01 - 0.6 6 rv/C Free molecules 7 0.01 6 r2v/1.66 1 which is used to define these regimes is the Knudsen number, Kn = 1/Dp = mean free path/particle diameter The flow regimes as defined by the Knudsen number are listed in Table 1, which also contains expressions for the drag force for each of the regimes. Table 1 Flow Regimes Defined by the Knudsen Number Regime Kn range Drag Force Dp range, m Continuum 0 - 0.1 0.6 6 rv Transition 0.1 - 7 0.01 - 0.6 6 rv/C Free molecules 7 0.01 6 r2v/1.66 1 = 1.83x10-4 poise, kinematic viscosity of air; r, particle radius; v, particle velocity, 25 C, Cunningham slip correction factor. The molecular flow field in the vicinity of the particle is illustrated schematically in Fig. 1a, b and c, for the three flow regimes. Free Molecular Regime. In the free molecular regime where the mean free path is much greater than the particle diameter, the molecules that collide with the particle come from a far distance where their velocity distribution function is not perturbed by the presence of the particle. The velocity distribution of such molecules is Maxwellian, and hence the momentum exchange with the particle can be readily calculated by the known laws from kinetic theory of gases (Epstein 19 ). Epstein's expression for the drag force, FD in the free molecular flow is FD 4 2 r c nmv 3 Where is a factor characterizing the surface ( = 1.44 for diffuse reflection); r2 is the cross sectional area of the particle; n, m and <c> are the number, mass and mean speed of gas molecules impinging on particles. Here we should note that the gas-particle collision frequency is proportional to the cross sectional area of the particle and the number concentration of the gas molecules. In this regime, the particles behave like giant molecules, subject to the laws of the kinetic gas theory. Continuum regime. For particles with Dp >>1 the surrounding gas medium appears as if it were a continuum. In this regime, the particles severely perturb the surrounding gas molecules, both by displacement and by changing their velocity field (Fig. 1c). The flow field can be calculated from classical fluid dynamics by solving the properly simplified NavierStokes equation (Schlichting, 1968). The simplifications arise from the fact that the inertial terms associated with the fluid displacement rate are small compared to the viscous terms and they can e omitted from the Navier-Stokes equation. The resulting expression for the drag force in "creeping " flow or inertialess flow is given by FD = 6 r v where is the air viscosity. This expression is frequently referred to as Stokes law and its main feature is that the drag force is proportional to the particle radius. The derivation of Stokes law also requires that the air is incompressible, particles are rigid spheres and that no interfering objects are in the vicinity of the particle. 26 Figure 11. Stokes law, which rests on the assumption of slow, creeping motion is only applicable CD 24 Re 4 1 (Re) 3 for Reynolds number less than one. For Re = (VDp/) > 1 inertial terms in the Navier- CD Fd 2 ( ρ p v )( 14 πD p ) 1 2 2 Stokes equation become important and subsequently boundary layer separation will occur. For an arbitrary flow regime, it is customary to express the drag in terms of the drag coefficient CD, where v is the relative particle velocity and p the air density. The value of CD depends on the Reynold number as shown in Figure 3. For Re 1 CD = 24/Re, and for 1 Re 400 an approximate formula holds within 2% of experimental results (Shlichting, 1964). For atmospheric aerosols, with Dp < 100 m, the characteristic velocities are such that the Stokes law is seldom violated. 27 Transition regime. The gap between the free molecular and continuum regime is characterized by particle size which is comparable to the mean free path of the suspending gas. In this regime, the particle perturbs sufficiently the velocity distribution function of the surrounding gas particles, such that the Maxwellian velocity distribution function and kinetic theory of gases is no longer applicable. On the other hand , the molecular concentration in the particle vicinity is sufficiently low, such that rarefaction effects cannot be ignored and hence the continuum is not established. In principle, the complete molecular velocity field in the vicinity of the particle and the resulting drag force can be calculated by solving the Boltzman equation with proper boundary conditions (Hapman and Cowling, 1959). Unfortunately, complete solutions of the Boltzman equation for the entire transition regime is practically not feasible due to the difficulties arising from determining the "collision integral". Figure 12. There are several approximate methods for the determination of the drag force in the transition regime that are reviewed by Hidy and Brach, 1970. For our purposes here, it is convenient to utilize the empirical Cunningham slip correction factor, C, which is applied to the continuum regime to correct for rarification FD = (FD) continuum/C = (6 r v)/C 28 The Cunningham slip correction, C, is an empirical expression based on Millikan's data for oil droplets (Fuchs, 1964). C = 1 + 2.492 Kn + 0.84 Kn e-0.44/Kn For small Kn, C 1; For Kn>>1, C 3.333Kn. For PO = 1, Kn = 0.666/Dp. At pressure P, which is different from PO, the mean free path is l= PO/PO, hence Knudsen number is (Kn) p Po (Kn) Po P The effect of pressure on the flow regime is most notable in the upper atmosphere where the pressure is P = PO exp (-…H). For example, at the atmospheric tropopause, at H = 20 km, the air pressure is P = …atm. With mean free path l = …m. Hence, a 0.1 m particle with Kn = is in the near free molecular regime at the tropopause, while it is in the near continuum regime at sea level. The functional form of C is such that for Kn 0 and Kn , the corrected drag force assymtotically approaches the expressions valid in continuum and free molecular regimes, respectively. The value of C is graphed in Figure 2. Mobility, Relaxation Time, Stopping Distance. Generally, particles follow the mean motion of the air molecules, i.e. they travel along fluid streamlines. This is due to the intense suspension forces caused by molecular bombardment from all directions. If, however, external forces act on a particle, such a force field may cause a deviation from the fluid environment and a particle will acquire a velocity relative to the fluid. As we have noted before, any relative motion is counteracted by a drag force due to impinging molecules, directed opposite to the direction of motion. Mobility. The relative velocity of a particle due to a force field is determined by the strength of this force as well as by the properties of the particle and the suspending air. These include the particle size, shape, viscosity, density, and mean free path of air, etc. For particles in the Stokes regime, it may be stated that the particle relative velocity will be proportional to the force that causes the relative motion. The proportionality constant, B, is called the mobility. Vp = BF or B= V/F Mechanical Mobility is the proportionality constant between the particle velocity and the drag force, i.e. B = V/Fd = C/(6r) For Kn < 0.1, B r-1 and for Kn > 7, B r-2. Thus, small particles are more “mobile” than larger ones. Particles that carry positive or negative electric charges are subject to electrical forces. The external electrical force exerted on such a particle is F = npeE, where np is the 29 number of elementary charges on a particle (1, 2, 3, etc.), e is the electronic charge 4.8 x 10 –10 esu, and E is the electric field dV/dx (V/cm). For a particle which is in a steady state motion, the electric force equals the resistance force. Hence, npeE= And V = npeB/E = ZpE where Zp = (npeC)/(6r) Is the electrical mobility of a particle. As noted above, Zp is proportional to the mechanical mobility and also increases with increasing quantum units of electric charge per particle. The electrical mobility is a useful parameter for the description of aerosol removal in charged clouds and for the size distribution measurement by electrical mobility analyzers (e.g. Whitby et al., 1972). There are several other forces, including the “foretic” forces (Davies, 1967), which may cause a relative particle motion and for which appropriate mobilities may be defined. Relaxation time. Consider a particle projecting into quiescent air with the initial velocity VO. Due to the drag force, the particle velocity will diminish with time. The rate of change of velocity is dV stated by the equation of motion, which arises from the force balance. At any instant of time, the inertial force given by Newton’s law m(dV/dt), is balanced by the fluid resistance force: m (dV/dt) = -Fd = -V/B where m = 4/3 r3 p Is the particle mass. Integration with initial condition V = VO at t = 0 yields an expression for velocity decay, V = VO e –(t/mB) = VO e –(t/) where = mB is the relaxation time. A similar expression may be obtained if the air is suddenly accelerated to velocity VO, such that V – VO (l – e –(t/)) Hence, after a time lag, 3t, the particle will attain the fluid velocity. The relaxation time for a 10 m particle is ~300sec. Stopping distance. Consider again a particle projected into a quiescent air. After sufficient time, the particle will stop and the total distance traveled, the stopping distance, s, may be obtained by integrating the equation of motion one more time. s 0 0 Vdt V O e t dt VO The stopping distance for a 10m particle, for VO = 10 cm/sec is s = 30m. Thus we may immediately conclude that for atmospheric particles with Dp < 10m, the relaxation times and stopping distances are rather short. Thus it may be stated that any change in fluid velocity will cause an instantaneous response in the appropriate particle 30 velocity. In such a case, the particle velocity may be determined from the quasi steady state equation of motion, V BF u where F is the external force and u is the air velocity vector. Settling velocity. A particle in a gravitational or centrifugal force field will be accelerated until the drag force equates the external body force. The steady state velocity is called the terminal velocity or the settling velocity, VS. For gravitational settling, VS may be obtained by equating the gravitational force, Fg, with the drag force, Fd. In the Stokes law regime (4/3) r3 (p - a)g = 6 r VS where a is the air density and g the gravitational constant. Since a << p, the air density may be neglected. The settling velocity is then VS = (2 r2 p g C)/(9) The main feature of VS is that it is proportional to the square of the particle size. For example, for Dp = 10 m, VS = 0.5 cm/sec, while for 100 m, VS = 50 cm/sec. The settling velocity is also proportional to particle density. For large particles settling at high speed (Re > 1) (4/3) r3 p g C = ½ a VS2 r2 Cd 8rρpg VO 3 ρaCd 1 2 In this case, the drag coefficient, Cd, itself depends on VS and hence the calculation of VS requires a numerical solution, e.g. by successive approximation. This method is only required, however, for settling particles with Dp > 100 m (VS = 50 cm/sec). 2. AEROSOL OPTICS Aerosol optics is concerned with the propagation of visible or near visible electromagnetic radiation through aerosol clouds and with the processes associated with the radiation transfer. From the point of view of radiation transfer, and aerosol system consists of randomly suspended discrete mass centers, which interacts with the radiation by absorption and scattering. Radiation The Physics of Radiation Radiant energy may be alternatively envisioned as being transported either by electromagnetic waves or by photons. Neither point of view completely describes the nature of observed phenomena. Nevertheless, these separate concepts have considerable utility. For example, the scattering from a single particle within a radiation wavefront 31 may be predicted through use of electromagnetic theory, whereas quantum theory is employed in determining the properties of gaseous absorption or multiple scattering media. Radiation travels at the speed of light. Thus, from the viewpoint of electromagnetic theory, the waves travel at this speed. Alternatively, from a quantum point of view, energy is transported by photons, all of which travel at the speed of light. (This differs from molecular transport in that all the photons have the same speed.) There is, however, a distribution of energy among the photons. The energy associated with each photon is h, where h is Planck’s constant and is the frequency of the radiation. Each photon also possesses a momentum h/c, where c is the velocity of light within the medium through which the radiation travels. Three parameters may be employed in characterizing radiation. They are the frequency , the wavelength , and the wave or photon speed c. Of these, only two are independent, since they are related by c=. The choice as to whether to employ or as a characteristic parameter is somewhat arbitrary, although has one advantage in that it does not change when radiation travels from one medium to another. The speed of light c within a given medium is related to that in a vacuum, co, by c = co/m m = (cos)/(cos') m = n - in', where m is the index of refraction. In vacuum, the index of refraction is unity while in gases it is approximately one. For common liquids and solids it is between 1 and 3. The wavelength range encompassed in aerosol optics which is of concern to us is the visible and the near-visible radiation i.e. 0.3 < < 1.0 m. The various radiation subranges are illustrated in Fig. 1 Figure 13. Classifications of Radiation. Absorption and Scattering The basic problem of aerosol optics is the change in radiation due to its interaction with the suspended particulates. The effect of visible radiation i.e. radiation pressure, absorption, etc., on the particle behavior may in general be neglected. When a particle is irradiated with light of a given wavelength, two different physical processes will occur. The incoming radiation can be transformed into other forms of energy, such as heat, energy of chemical reactions, or radiation at a different wavelength. In such a case the 32 energy transformation is called absorption. The absorption of radiation is associated with transitions of the energy levels of the atoms or molecules that constitute the aerosol. Absorption terminates the path of a photon. In visible range very few gases, such as NO2, have spectral absorption bands. The most important gaseous absorption bands are in the ultraviolet and infrared regions. In addition to absorption, a medium may also scatter photons. Scattering is defined as any change in the direction of propagation of the photons. This process is physically due to local inhomogeneities within the medium, and such inhomogeneities may result from suspended solid particles or liquid droplets within the gas. In addition, scattering can also be produced by the gas molecules. When radiant energy is scattered with no change in frequency, the scattering is referred to as coherent scattering. If the scattering of radiation within a gas is strictly molecular scattering (i.e., there are no foreign particle present), it is designated as Rayleigh scattering. The Rayleigh theory predicts that the spectral intensity of the scattered radiation will vary as the fourth power of the frequency; that is, the scattering is predominantly at the shorter wavelengths. This accounts for the fact that the sky appears blue, for the preferential scattering in the atmosphere involves the short wavelength blue light. This is also the reason why sunsets are red, for the long wavelength red light suffers less attenuation in traversing the large atmospheric path length. Although Rayleigh scattering is an important mechanism in global atmospheric phenomena, it is usually unimportant for microclimatological applications due to the short path lengths involved in the latter. Scattering can, however, play an important role in radiation energy transfer when foreign particles are present. Typical examples include clouds, fogs, and air pollution particulates. In these cases scattering may encompass the combined and interactive effects of reflection, refraction, and diffraction. A theory that is pertinent to such situations is Mie scattering, which is concerned with electromagnetic scattering from spherical particles with sizes on the order of the wavelength of the incoming radiation. The foregoing was a discussion of the physics of absorption and scattering. Attention will now be turned to the formulation of the processes of absorption and scattering in terms of defined radiation properties. Intensity of Radiation To characterize the amount of radiant energy that departs from a surface along a certain path, the concept of a single ray is inadequate. The amount of energy passing in a given direction is described in terms of the intensity of radiation, which is denoted by i. With reference to Fig. 2, the intensity of radiation is the radiant energy leaving a surface per unit area normal to the pencil or rays, per unit solid angle and per unit time, where the differential solid angle is d. To illustrate the use of intensity, let d represent the radiant energy per unit time and unit area leaving a given surface in the direction and contained within a solid angle d. 33 Figure 14. The intensity of radiation. Then clearly i dΦ dω cosθ (1) The energy flux passing from the surface into the hemispherical space above the surface is then obtained by integration Φ i cosθ dω Δ (2) where the symbol denotes integration with respect to solid angle over an entire hemisphere. As illustrated in Fig. 3, the differential solid angle d may be expressed in terms of the angles and of a spherical coordinate system centered on the surface. Upon recalling that the differential Figure 15. Integration of intensity over solid angle. 34 solid angle is the surface element on the hemisphere divided by the square of the hemisphere radius, it follows that d = sin d d(area on a unit sphere). Consequently, integration over the entire hemisphere yields Φ 2π π 2 0 0 i cosθ sin θ dθ d (3) There are several instances in which the intensity of radiation is independent of direction. For such cases equation (3) reduces to = i (4) In the foregoing definition of the intensity i, specific reference has been made to radiation leaving a surface. When dealing with the radiation concerned with the net rate at which energy is locally transferred within the medium. The intensity of radiation will, in this instance, be designated by the symbol I and defined as the local net transfer of radiant energy per unit area normal to the pencil of rays, per unit solid angle, and per unit time. Although the foregoing discussions of intensity have been concerned with the total energy (over all wavelengths), the formulations apply on a monochromatic basis by appending the subscript . Thus the spectral (monochromatic) intensities are i and I, while the expressions i id I 0 I d 0 (5) Relate the total and spectral quantities. Absorption, Scattering, Extinction Consider a monochromatic beam of radiation with intensity I, as illustrated in Fig. 4. As the beam traverses a path length ds it will undergo partial attenuation as the result of local absorption within the medium. The amount of absorption is assumed to be directly proportional to both the thickness ds and the incident intensity I. Thus the amount of monochromatic absorption per unit time, per unit area normal to the pencil of rays, and per unit solid angle may be written as kIds (6) 35 Figure 16. Absorption and scattering of incident radiation. The constant of proportionality k is the monochromatic absorption coefficient (this is actually a volumetric absorption coefficient, as distinct from the often-employed mass absorption coefficient.) for radiation of wavelength . Since the volume per unit surface area of the crosshatched element in Fig. 16 is ds, the monochromatic absorption within the medium per unit time, per unit volume, and per unit solid angle is kI (7) By integration of equation (2) over all possible values of the solid angle ( = 4), the local monochromatic absorption per unit time and per unit volume within the medium due to all incident beams is expressed as k Id 4 (8) where it has been assumed that the medium is isotropic (i.e., k is independent of direction). A parallel development can be carried out for scattering. Again, with reference to Fig. 16, the monochromatic beam of intensity I will further be attenuated due to scattering. The monochromatic energy that is scattered per unit of time, per unit area normal to the pencil of rays, and per unit solid angle may be characterized as I d (9) where is the monochromatic scattering coefficient. Correspondingly, on a unit volume basis, the expressions γλ I λ I (10) dω 4π (11) 36 respectively characterize the scattered energy due to a single incident beam and to all incident beams. The analogous description of radiation attenuation due both to absorption and to scattering leads to the following representation of the combined effects of both processes k I ds + I ds = I ds (12) where = k+ is the monochromatic extinction coefficient. The parallelism between absorption and scattering ceases, however, at this point. Once photons have been absorbed, no further consideration need be given to them. On the other hand, scattered photons continue to transport energy throughout the medium, and these must be taken into account. To facilitate the description of the radiant energy flux leaving a volume element as the result of scattering of incident radiation, consider the schematic representation as shown in Fig. 17, where s is a scalar distance measured along a pencil of rays. A portion of the incident beam I (s, ’, ’) will be scattered by the crosshatched volume element in accord with equation (3). It will be assumed that the scattering is coherent. Figure 17. Scattered energy leaving a volume element. The directional distribution of the scattered energy is characterized by the scattering function P (’, ’: ,) such that P (’, ’: ,) (d / 4) (13) represents the probability that radiation incident in the direction (’, ’) will be scattered in the direction (,). Since the magnitude of the scattered radiation corresponding to the incident beam I (s, ’, ’) is described by equation (3.4), the monochromatic energy per unit time, per unit solid angle, and per unit area normal to I (s, ’, ’) that is scattered in the direction (,) is I (s, ’, ’) P (’, ’: ,) (d / 4)ds (14) 37 The integral of the foregoing quantity over all values of the solid angle must coincide with equation (3). From this, it follows that 1 4 P ( ' , ': , )d 1 4 (15) which is consistent with the definition of P as a probability. An often-used (though not necessarily often-realistic) assumption is that of isotropic scattering. In this case, radiation is scattered uniformly in all directions so that P (’, ’: ,) = 1 (16) Scattering and Absorption by Single Particles In this section we shall consider the scattering characteristics of single particles and provide the information required for the calculation of radiative transfer and visibility in the hazy atmosphere. The problem is to relate the properties of the scatterer, its shape, size and refractive index, to the intensity and angular distribution of the scattered light. It is of interest to note that scattering is not restricted to the visible part of the electromagnetic radiation spectrum and that the waves by satellites, the scattering of microwaves by raindrops, scattering of thermal radiation by cloud droplets, scattering of light by small particles, and electron scattering by molecules are similar phenomena, since in each case the wavelength is of the same magnitude as that of the scatter. Thus, as a natural scaling fact or for scattering is the wavelength of the incoming radiation and it is used in the dimensionless parameter, called the optical size parameter, . = 2 r/ where r is characteristic particle size; In the case of a sphere it is the radius. In the following discussion, the incident radiation will be assumed to be monochromatic. Role of Particle Size The primary role of particle size is that it determines the regime of interaction with radiation. For particles much smaller than the wavelength << 1, the presence of the particle does not perturb the wavefront of the electromagnetic radiation. The physical process of scattering in such a regime is described by dipole interaction with electromagnetic radiation and it is also referred to as Rayleigh scattering. When the particle size becomes comparable to the wavelength, then the wavefront suffers a substantial distraction in the vicinity of the particle. Waves scattered by a particle will tend to interfere positively and negatively with the surrounding radiation wavefront, which leads to a peculiar interference pattern in the vicinity range and particles in this size range are called Mie particles. For a particle with diameter at least ten times the wavelength, the interaction with the wavefront again becomes relatively simple. Such a particle will reflect, refract (the lens effect), and diffract certain amounts of radiation. These particles fall in the regime of geometrical optics or ray tracing. 38 Beyond these qualitative changes in the radiation interaction regime, the particle size has a crucial role in determining the amount of light each particle will scatter. Cross Section and Efficiency When the electromagnetic wavefront passes a particle, the radiation in the vicinity of the particle will be abstracted from its original path, by scattering and absorption. This fraction of incoming radiation is extinct from the original beam. Hence, the extinction cross section of a particle, Cext, is the area in the particle where all the incoming radiation changes directions or gets absorbed. The scattering cross section Cscat and the absorption cross section, Cabs, are defined analogously and follow the relation: Cext = Cscat + Cabs Intuitively, the extinction cross section is the cross sectional area of the particle r2, i.e. all the energy impinging upon the particle cross sectional area is extinct with unit efficiency factor. However, in reality this efficiency factor Qext, is either less than or greater than unity depending on and the refractive index m. Hence, Cext = Qext r2 The efficiency factors for scattering, Qscat, and absorption, Qabs, are defined in a similar manner. In analogy of the cross sections, it also holds that Qext = Qscat + Qabs One would expect that for very large spheres, certainly for large absorbing spheres, the extinction cross section would approach the physical cross section i.e. Qext ~ 1. Paradoxically, this is not so. A large scattering sphere extincts twice its geometrical cross section. The cause for this discrepancy is explained by the “diffraction paradox.” Efficiency per Unit Mass It is often desirable to define and compute the efficiency of extinction per unit mass of aerosol. The mass extinction efficiency Mext is defined for spherical particles as M ext C ext C Q 4ext 3 ext1 M ρ p 3 πr ρp 3 r The mass scattering and absorption efficiency factors are defined analogously and satisfy the relationship Mext = Mscat + Mabs The utility of the mass efficiency factors lies in that they are identically potency functions (see pp (8) for scattering and absorption). Albedo for Single Scatter In the general case, the photons may be scattered and absorbed by the particles, thus only a fraction of the incident photons will be leaving the particle. The albedo for single scatter, , is defined as the fraction of light lost from the incident pencil due to scattering, while ( 1 – ) represents the absorbed fraction of the energy. Thus, the albedo is defined as 39 = Cscat / Cext = Qscat / Qext = Mscat/ Mext The single scatter albedo is particularly useful in multiple scattering calculations. Scattering Diagram and Phase Function, Asymmetry Factor The scattered wave of any point in the distant field has the character of a spherical wave in which energy flows outward from the particle. The direction of scattering is characterized by the scattering angle and azimuth angle . The most important property of the incident and scattered wave is the intensity, I. The intensity of electromagnetic radiation is the rate of energy flow across a unit area (erg/cm2sec) perpendicular to the direction of propagation. In optics this is also called irradiance. The intensity is occasionally also referred to as illuminance, i.e. luminous flux per unit area (lumens/m2 = lux). Both the incident and scattered waves are unidirectional, i.e. each confined to a narrow solid angle. The term I is the total energy flux in this narrow solid angle d. The waves are also assumed to be monochromatics, i.e. confined to a narrow frequency interval. Here we should recall that it is advantageous to use frequency rather than wavelength, since is independent of the refractive index of the medium. However, for practical reasons, the intensity is commonly defined in terms of wavelength increment d. If Io and I are the intensities of incident and scattered light respectively, and R is the distance from the particle, then I must be proportional to Io and R –2, and we may write I λ2 F( , ) 4π 2 R 2 Figure 18. The total amount of incident energy that changes direction of propagation is the amount confined to the scattering cross-section, Cscat. The same energy is now distributed in all directions as given by the dimensionless angular function called the scattering diagram, F (, ). From conservation of energy, we get Cscat 1 k 2 F(θ, )d 4 40 Wave number k = 2 / where d = sin d d is the infinitesimal solid angle and the integral is taken over all directions, 0 < < ; 0 < < 2. The scattering diagram, when divided by Cscat k2 yields the phase function P (,): 1 4 P(θ, ) 1 4 P (,) = F (,) / Cscat k2 which is the fraction of the total amount of light scattered into direction 2 and . The integral of P (,) over all directions is unity.Hence the phase function represents the probability of scattering in any given direction. The probabilistic interpretation of P (, ) will be useful for the photon tracing in multiple scattering problems. The value of P (, ) also depends on the size parameter, and the refractive index m. For spherical particles in natural nonpolarized light, the light scattering phase function has circular symmetry coinciding with the axis of incident beam propagation such that P(,) = P() Generally, the forward scattering (with > goo) dominates the backscattering and this asymmetry is quantified by the asymmetry parameter cos , defined as cos θ 1 4 P(θ, )cos θ d 4 The asymmetry factor increases with increasing particle size. Polarization Scalar waves, like sound, are fully described by the intensity. However, neither the incident light nor the scattered light is completely characterized by intensity. The transverse nature of light waves allows the phenomenon of polarization to occur. The additional parameters required for the full description are polarization and phase. Light consists of many simple waves with frequency 1014 sec –1 and with the duration of coherent wave trains 10 –8 sec. The simple waves are all monochromatic and completely (elliptically) polarized. Light that is commonly measured is the net effect of many simple waves and in general is partially polarized. Natural light, such as direct sunlight, is a mixture of uncorraleted simple waves. Over a time period of usual measurements ( t > 10 –8 sec) the electric vector exhibits no preferential vibration i.e. it is unpolarized. An arbitrary beam of light, of intensity I, consists of an unpolarized part and a totally polarized part. I = Iunpol + I pol The degree of polarization is defined as Ipol / I. The polarized part of the beam is in general elliptically polarized, and it can be further separated into a linearly polarized part, Ilp, and a circularly polarized part, Icp, where Ipol = ( Ilp2 + Icp2 ) (1/2). 41 The intensity of incident radiation complete with polarization may be represented by Stockes vector II = { Io, Qo, Uo, Vo}, where the quantities in the bracket are the four Stockes parameters. The scattered intensity at distance R in the far field (i.e. R >> ) is II = (1 / k2 R2) F IIo where the quantity describing the single scattering properties of scattering medium is F, the transformation matrix, four columns, and four rows of dimensionless numbers. For unpolarized natural incident light, the Stockes vector simplifies to Io = { Io, 0, 0, 0}. Hence, the intensity is adequate to describe this type of radiation. In the framework of the following discussion, we shall confine the discussion to unpolarized incident light. Unfortunately we cannot make that simplification for the radiation leaving the particle. The scattered radiation is in general polarized. For the present discussion it will suffice to say that the scattered intensity will have components vibrating perpendicularly (i 1) and parallel (i2) to the plane through the directions of propagation of the incident and scattered beams. The degree of polarization is (i1 – i2)/(i1 + i2). The scattered wave in the distant field from the particle is a spherical outgoing wave with an amplitude, S(), inversely proportional to distance. Any point in space is traversed by two wave systems, the incident and the scattered wave. The intensity of the scattered wave for perpendicular polarization is I = (i1 / k2R2) Io for parallel polarization I = (i2 / k2R2) Io and for incident natural light I = ((1/2) (i1 + i2) / k2R2) Io The intensity functions are related to the amplitude functions by i1 = | S1 () |2 i2 = | S2 () |2 The wave phenomena involved in light scattering are illustrated schematically in Fig. 19. 42 Figure 19. Schematics of the spherical wave of linearly polarized light leaving the particle. Much of the light scattering theory is concerned with the understanding of the intricate interference relationship between the scattered (diffracted, refracted, and reflected) and the incident light. In the next section, consideration is given to the basic physical processes, which cause light scattering to occur. Diffraction, Refraction, and Reflection The physics of light scattering by small particles can be described by the combined and mutually interfering effects of diffraction, refraction, and reflection. Reflection is a well known phenomenon and generally contributes < 5% of the scattering. Therefore, no further consideration will be given to its role at this point. Role of Diffraction Diffraction is an edge effect. It arises from the incompleteness of the wavefront passing the sphere. Figure 20. 43 A fraction of the radiation near the edge of the particle is bent towards the particle to “fill its shadow.” On rigorous physical grounds it can be shown that the amount of radiation diffracted is precisely the cross sectional area of the particle. This holds for any shape particle. Hence, the extinction cross section of a large sphere consists of the refracted (or absorbed) cross section, r2, and of the diffracted cross section, which is also r2. This explains the “diffraction paradox” that Qext = 2 (Van de Hulst, 1957) for large particles, and the polarization of diffracted light is the same as that of the incident light. Role of Refractive Index The refractive index plays three major roles in extinction: 1) phase shift, due to difference of speed of wave propagation inside and outside the particle, 2) bending of the wavefront direction (the lens effect), 3) dissipation (absorption). The speed of a wavefront entering the particle with n > 1 is reduced by the amount c = co / n, where co is the speed of light in a vacuum. Since the frequency remains constant, this leads to reduction of the wavelength within the particle. Consequently a phase shift develops between the wave within and outside the particle. Hence, the wave leaving the particle (i.e. center of the particle) may be in or out of the phase with its surroundings. This then may lead to positive or negative interference. We shall note later that it should also be stressed that phase shift is a necessary condition for scattering to occur, i.e. the refractive index difference at the particle interface needs to be large. This is satisfied when 2(m-1) > 1. (van de Hulst, 1957). Figure 21. Phase shift. The refraction index is also responsible for bending the arriving wavefront in a similar manner as observed in the well known convex lens effect. For small refractive index, the bending is weak, while at high n, the dispersion is strong as illustrated schematically in Fig. 22. The refracted light from a particle contributes most of the polarization. 44 Figure 22. The "lease effect" of refractive index For absorbing media such as metal oxides and elemental carbon, the refractive index, m, is complex. The refracted wave now becomes inhomogeneous, with decaying intensity (amplitude) as it penetrates the particle, Fig. 23 Figure 23. The refractive index, m, is now expressed as a complex number m = n – in’ where n and n’ are the real and imaginary parts of m. The radiation intensity in the particle decreases by exp (-4 n’) , and the amplitude by exp (-4 n’). Refractive Index for Shortwave Radiation The value of the refractive index for dry atmospheric aerosol particles ranges from 1.4 < n < 1.6. For water, the value is 1.33. For a hygroscopic or deliquescent particle, the refractive index will approach that of water as the particle grows at high humidities. Since the growths of particles with humidity is only known for a few pure substances, the RH dependence of refractive index of atmospheric aerosols is not predictable on theoretical grounds. Barnhardt & Strate (1970) quote an empirical formula N = 1.54 + 0.03 ln (1-RH/100). Hanel (1971) proposed a linear interpolation between the dry particle refractive index and water, based on the volume fraction of the dissolved matter. The refractive indices of crystalline aerosol materials have been compiled by Bullrich (1964) and are shown in Table 1. These crystalline compounds have 1.48 < n < 1.64 in 45 dry state and all f them are transparent in the visible window (i.e. the imaginary part of their refractive index is negligible). The imaginary part of the refractive index is responsible for the absorption (heating) of particles, but its value for atmospheric aerosols is not well established. For urban aerosols it is considered to range between 0.005 to 0.03. Absorption in the visible region may be due to organic aerosols or metal (ferrous) oxide, but is believed to be dominated by carbon (soot) particles emitted from combustion sources. Twetty & Weinman (1971) have reviewed the real and imaginary refractive indices of graphites, soot, coals, etc., and gave an average value of m=1.8-0.5 i. Bergstrom (1972) proposed 0.63 < n' < 0.69 in the visible window, but he also stressed the wavelength dependence of both n and n'. Bergstrom's n, n' values for the 0.3-3m range are given in Table 2. Figure 24. Real (nr) and imaginary (ni) parts of the refractive index of quartz (silica). (After Conel, 1969). 46 Figure 25. Particle Refractive Index Particle Refractive Index NH4Cl 1.64 CaSO4 1.57 NH4NO3 1.60 KCl 1.49 (NH4)2SO4 1.52 Na2SO4 1.48 MgCl2 1.54 SiO2 1.49 NaNO4 1.59 K2SO4 1.49 Table 1. The refractive index of atmospheric aerosol materials which have a crystalline structure (Bullrich, 1964). The calculation of the effective refractive index of a mixture of particles such as the atmospheric aerosol, is possible if the size, chemical composition distribution function is known. In that case, the average real and imaginary refractive indices may be extracted from properly calculated scattering and extinction coefficients. The real and imaginary parts of the refractive index of atmospheric aerosols have been examined by a variety of experimental techniques, some requiring aerosol deposition on filter. In other methods, n and n’ were inferred from in situ optical measurements. Table 3 summarizes n and n’ data for soot, fly ash, atmospheric aerosol, and dry dust. Prudent users of such tables will carefully compare their own application with the conditions that the above data were taken. 47 Refractive Index for IR Radiation In the wavelength range 0.7 < < 14 m, the real and imaginary refractive indices of materials cannot be taken as constant. Most materials exhibit absorption bands particularly in the 1-10 m region, in which n’ will vary drastically. If this variation of n’ occurs in the “transparency window” of the atmosphere (8-14 m) the analysis needs to be with care (in spectral bonds). The extinction is generally dominated by particle sizes comparable to the wavelength of radiation. Hence, for the transparency window, the particle sizes 8-14 m are of particular importance, which is dominated by earth’s crust materials: silicates and metal oxides. Conel (1969) reviewed the optical properties of silicates and gave n and n’ (Fig. 24) from measurements of Spitzer & Kleinman (1961). The imaginary part of refractive index for propane soot, chimney soot, soot from precipitation dust, coarse dust, and fine dust, has been measured (Fig. 25) by Volz (1972) over a 2-30 m wavelength range. While soot is an efficient absorber, dust has been shown to be quite absorbing (n’ ~ 0.1) in the 8-14 m band. Further discussion on the subject is given by Paltridge and Platt (1976, pp 287). MIE Solution In principle, the interaction of the particle with the incoming electromagnetic radiation may be determined by solving Maxwell’s equations for the boundary conditions corresponding to the particle shape. Solutions, however, are available only for few shapes: spheres, spherical shells, ellipsoids, and infinite cylinders (van de Hulst, 1957). The first solution of the Maxwell equations for dielectric (transparent) and absorbing spheres was given by Gustav Mie (1908). The equation of concern is the wave equation: + ( 2m / )2 = 0 which is satisfied by both the electric and magnetic field strength. This equation is solved for within and outside the particle, such that the solutions are matched at the particle boundary. The method of separation of variables yields a formal solution that may be worked out with the use of Bessel functions. Mie’s solutions of the spherical wave equation for a homogeneous sphere are given in terms of the intensity functions i1 and i2, vibrating perpendicularly and parallel through the plane of incident wave propagation and scattering direction. i1 S1 2 2n 1 a n n b n τ n n 1 m(n 1) 2n 1 i 2 S2 ( ) (a n τ n b n n ) n 1 n(n 1) 2 2 2 48 The polarized intensity functions are computed from the above infinite series where n and n are angular functions derived from Legendre polynomials. polynomials are easily computed from recursion relations. The first two are: 1 () = 1 2 () = 3 cos () (m) Legendre 1 () = cos 2 () = 3 cos 2 n2 k2 0.3 1.84 0.70 0.4 1.88 0.69 0.5 1.94 0.66 0.6 1.99 0.64 0.7 2.03 0.63 0.8 2.07 0.61 0.9 2.09 0.60 1.0 2.12 0.59 1.5 2.14 0.65 2.0 2.17 0.75 2.5 2.21 0.86 3.0 2.26 0.98 0.42 1.41 2.53 0.589 1.79 3.33 0.75 2.19 4.36 1.0 2.63 5.26 2.25 3.95 9.20 (a) Carbon (b) Nickel Table 2. Indices of refraction (2 = n2 - ik2) for carbon and nickel at various wavelengths. The actual index of refraction of quartz in the solar spectrum is about = 1.52--0.0 I (Peterson and Weinman, 1969); however, the value of 1.50 was used to be consistent with that comomnly employed for the natural aerosol (Quenzel, 1970; Yamamoto and Tanaka, 1969; and others). 49 ni 1.8 1.55 nr Type Twitty and Weinman (1971) 0.5 soot and coal dust Grams et al. (1972) 0.044 fly ash Fischer (1970) 0.01 atmospheric aerosol Ivley and Popova (1973) 1.65 0.005 atmospheric aerosol Lin et al. (1973) 0.01 atmospheric aerosol Lindberg and Laude (1974) 0.007 dry land dust Table 3. Values of the real (nr) and imaginary (ni) parts of the refractive index of atmospheric particles for the short-wave region observed from various experimental measurements. (Paltridge and Pratt, 1976). Figure 26. The two sets of functions of the scatttering angle , which occur in the Mie formulae, for n = 1 to 6. (van de Hulst, 1957). The heart of the Mie scattering problem is the computation of the series coefficients an and bn, arising from separation of variables. These functions depend on the size parameter, , and the complex refractive index m = n – in’. The expressions an and bn involve Bessel functions (van de Hulst, 1957) and can also be computed with recursion relations. Appropriate methods of numerical computation have been discussed by Kattawar & Plass (1967) and Dave (1969). 50 The scattering and extinction efficiency factors also follow from the coefficients an, bn (van de Hulst, 1957). as does the asymmetry factor. 4 cosα 2 x Q scat n(n 2) 2n 1 Re(a n a *n 1 b n b *n 1 ) Re(a n b *n ) n(n 1) n 1 n 1 The values of an and bn rapidly approach zero when n . Therefore, the total number of terms that should be calculated in these series (Hansen & Travis, 1974) is n . The infinite series can be physically interpreted as a multipole expansion of the scattered light (Mie, 1908). The coefficients a1, a2, and a3, i.e. specify the amount of electric dipole, quadrupole, and octupole, while the bn are coefficients of magnetic multipole radiation. For particles which are small and do not have large refractive index, only the electric dipole radiation is significant and the well-known Rayleigh scattering results. Size Distribution. In nature a distribution of particle sizes is usually encountered. The scattering and extinction coefficients are r2 r2 r1 r1 γ C scat (r)n(r)dr πr 2 Q scat (r)n(r)dr r2 r2 r1 r1 C ext (r)n(r)dr πr 2 Q ext (r)n(r)dr (2.48) where n(r)dr is the number of particles per unit volume with radius between r and r + dr, r1 and r2 are the smallest and largest particles in the size distribution, and under the assumption of independent scattering the transformation matrix for a unit volume is given by r2 F ( ) F ij ( , r) n(r) dr ij r1 51 (2.47) Fij(, r) is one of the elements of the transformation matrix for a particle of radius r. The normalized (dimensionless) phase matrix follows from P ij ( ) 4 ij F ( ) k 2 (2.49) And the single scattering albedo from (2.50) It is straightforward to make computations for any size distribution of spheres. However it is important to have systematic computations which allow an understanding of the effect of the size distribution. This is required if measurements are to be inverted to yield properties of scattering particles. To facilitate inversion of radiation measurements the size distribution must be described with the minimum number of parameters. Clearly the first parameter should be some measure of the mean particle size. The arithmetic mean is r2 r rn(r)dr r 1 2 r n(r)dr N r1 r1 r2 n(r)dr r1 (2.51) Where N is the total number of particles per unit volume. But since each particle scatters an amount of light proportional to Cscat = r2 Qscat, the "best" single parameter describing the scattered light is the mean radius for scattering. r2 rscat r π r 2 Q scat ( , n r , n i ' ) n(r)dr r1 r2 π r 2 Q scat ( , n r , n i ' ) n(r)dr r1 (2.52) The appearance of Qscat (, nr, ni') in (2.52) makes rscat an inconvenient parameter. However, if rscat is larger than the wavelength a parameter which is almost as "good" can be obtained by omitting Qscat in (2.52). Thus we define the effective radius as r2 reff r π r r1 r2 2 n(r)dr 2 π r n(r)dr r1 r 1 2 r π r 2 n(r)dr G r1 52 (2.53) Where G is the geometric cross-sectional area of particles per unit volume. Similarly, as a measure of the width of the size distribution, we define the effective variance, v eff 1 Gr eff2 r2 (r - r eff ) 2 π r 2 n(r)dr r1 (2.54) Where reff2 in the denominator makes veff dimensionless. As a measure of the departure of the distribution from symmetry we define the effective skewness, s eff r2 1 3 eff Gr v (r - r 3/2 eff r1 eff ) 3 π r 2 n(r)dr (2.55) These definitions are analogous to characteristics used in statistics to describe frequency distribution (e.g., Kendall and Stuart, 1963), with r2n(r) corresponding to the frequency distribution. It is useful to have a standard analytic size distribution for theoretical computations. We employ the distribution used by Hansen (1971b), n(r) constant r (1 3b)/b e r/ab (2.56) As a standard distribution, because it has the simple properties a = reff b = veff for the size distribution (2.56) (2.57) As may be verified by substitution into (2.53) and (2.54) with r1 = 0 and r2 = . The standard distribution (2.56) is a variation of the gamma distribution; other forms of the gamma distribution have been used extensively for cloud particles, e.g., by Khrgian (1961) and Deirmendjian (1964). Figure 29 illustrates the standard distribution for several values of a and b. This distribution, including the normalization constant, is defined for 0 b = veff< 0.5 and has an effective skewness seff = 2b. Larger values of veff can be obtained by adding a third parameter to (2.56) or by using the log-normal or power-law described in the following subsection. Numerical Examples. Mie scattering computations are a simple task for modern computers and there now exist several books and reports containing extensive tables of numerical results. Nevertheless it is useful to have graphical examples which clearly illustrate the effect of the size distribution and refractive index on the scattered light. 53 The first graphs which we show are for quantities independent of the scattering angle, Qscat, , and cos. Qscat, the efficiency factor for scattering, is defined for a size distribution as r2 Q scat G π r 2 Q scat (r) n(r)dr r1 r2 π r 2 n(r)dr r1 (2.58) Scattering Efficiency Factor, Qscat The value of Qscat as a function of and refractive index m is shown in the tridimensional plot, Fig. 27. At m = 0, the value of Qscat is zero. This is due to the observation made earlier, that in order for scattering (phase shift) to take place, a difference in m at the particle interface is necessary. The Qscat curves for m > 1 are characterized by a series of major maxima and minima. The major maxima and minima are due to interference of waves diffracted and transmitted by the particle. Since the phase log, 2(n-1), of transmitted light is greater at high n, increasing n (Fig. 27). At large n, the amplitude of the interference oscillations also increases. Figure 27. Three-dimensional view of smoothed values of Qsca as a function of and ml. The grid lines designate constant values of . (Kerker, 1969) The behavior of Qscat is shown in more detail in Fig. 28. In addition to the major oscillations, a ripple on the Qscat curve also arises from the last few significant terms in the Mie series. Physically the ripple arises from light waves grazing the sphere. These rays set up surface electromagnetic waves, which travel around the sphere spewing off energy in all directions. At = 180o, when the surface waves splash and they give rise to enhanced back scattering intensity. The ripple is due to the interference of the diffracted and surface waves (van der Hulst, 1957). Fig. 28 also shows Mie computations of Qscat for polydispersed aerosols with varying width. The size distributions having different 54 spread factor, b, are shown in Fig. 29. It is demonstrated (Fig. 28) that even slightly polydispersed aerosols do not exhibit the secondary interference peaks and ripples. However, the first maximum persists even for large values of b. Figure 28. Efficiency factor for scattering, Qscat as a function of the effective size parameter, 2a/. The standard size distribution (2.56) was used with four values of the effective variance b. For the case b = 0, 2a/ = 2r/ x. The refractive index is nr = 1.33, ni = 0. (Hansen and Travis, 1974) 55 Figure 29. Standard size distribution (2.56) for 2 values of a and three values of b. The size distribution is normalized so that the integral over all sizes is N = 1. (Hansen and Travis, 1974) For absorbing particles, the major maxima and minima and ripples are damped. For large particles (>> 1) the intensity of light waves traversing the diameter of the sphere is decreased by a factor exp (-4 n’), and thus the major maxima and minima are significantly damped for 4 n’ 1. 56 Figure 30. Efficiency factor for scattering Qsca, as a function of the size parameter x 2r/. The refractive index nr 1.33, with results shown for four values of m. Mass Extinction Coefficients The efficiency of scattering and absorption for unit mass of aerosol is a useful parameter for the estimation of the contributions of various sources to the total light scattering and absorption. The light extinction coefficient, , of a polydispersed aerosol population, consisting of a mixture of i different species is the sum of the contributions of all species. r2 M ext,i (r, n, n' )mi (r)dr i r1 where Mext,i (r, n, n’) is the mass extinction efficiency factor (see p. 60) for species I having real and imaginary refractive indices n and n’. The mass distribution with size of species I is defined on p. 10 and an example is shown on p. 21b and 38a. 57 A set of calculations for the efficiency factors per unit aerosol volume are shown in Fig. 19 (Patterson and Wagman, 1976). Another set of computations was performed by Foxwg (1976), who showed that the commonly found aerosol species have quite different Mext. While the absorbers, Fe and C have their peak at 1, the nonabsorbers H2O and SiO2 scatter most efficiently at 5. The mass scattering and absorption efficiencies were computed by Bergstrom (1973) for a set of real and imaginary refractive indices. His results for a wavelength of 0.55 m and indices of refraction of 1.5-0.02i, 1.5-0.05i, and 2-0.66i (carbon) are shown in Fig. 33. For nonabsorbing particles with n = 1.5, the scattering per unit mass peaks at r 0.4 m (Dp 0.8 m). For very small particles, i.e. Dp < 0.1 m the mass scattering function approaches Dp3 , and the efficiency curves become very small. For large particles Mext ~ r-1 as Qext approaches 2. For absorbing particles with << 1, the extinction is equal to the absorption and the absorption per unit mass is constant. Such particles will behave as gaseous absorbers, Bergstrom (1973). Bergstrom also performed a set of calculations for the IR wavelength of 9 m and n’ up to 2.0. In this spectral region the small particle limit extends to about Dp = 1.0 m. In this regime, r < 1.0 m, the mass absorption coefficient does not depend on the size distribution, which was also noted by Waggoner et. al. (1973). The increase in the small particle absorption is almost linear with increasing n’ up to n’ = 1.5. Figure 31. Light scattering per unit volume (100 of aerosol. Note that for = 3 3 3 3 1gr/cm , 1m /cm = 1g/m . The light scattering/volume is ~ double for m = 1.65 compared to m = 1.33. However, the density of materials at m = 1.65 is perhaps double that of water (1.33). Hence, the scattering/ mass is to be calculated as shown in Fig 32. m3/cm3) 58 Figure 32. Optical scattering cross section per unit mass ofr single particles of various materials at wavelengths at wavelenght of 0.6328 m. (The refractive indices and densities used in the calculations are: iron (m = 1.51-1.63i p = 7.86); carbon (m = 2.5-0.75i, p = 2.25); water (m = 1.33, p = 1.0); and silicon dioxide (m = 1.55, p = 2.66).4 The wavelength used is = 0.6328 m. (Foxwog 1975)). 59 Figure 33. Extinction and absorption coefficients per unit mass as a function of particle radius for four different refractive indices at a solar wavelength of 0.55 m. Figure 34. Absorption coefficient per unit mass of a log-Gaussian size 60 distribution of carbon particles asa function of the raidus of maximum concentration r0 for different values of the spread parameter A at a wavelength of 0.55 m. The mass absorption efficiency for a polydispersed carbon aerosol distribution is shown in Fig. 34. For carbon the efficiency is practically constant from small sizes up to Dp = 0.2 m, and then it falls linearly with increasing size. The numerical studies of Bergstrom shown above have clearly pointed to the importance of small particles in absorption, even for sizes below the “optical window,” 0.2-0.7 m. If we note that most of the elemental carbon is emitted from combustion sources and is concentrated in the 0.05-0.3 m size range, then these particles obviously deserve most of the attention when studying the absorption and atmospheric heating effects. Asymmetry Parameter, cos This parameter is a measure of the skewness of the phase function toward forward scattering. A symmetric phase function such as that of Rayleigh scattering has an asymmetric parameter, cos = 0, as shown in Fig. 35. On the other hand, for large optical size parameter, cos approaches the result for the geometrical optics phase function, cos = 0.87 (van de Hulst, 1957). In between it has a major peak and smaller oscillations for the same reason as the oscillations of Qext: phase lag in the particle and interferences with diffracted waves. An increase in the width of the size distributions tends to smooth out these fluctuations. The asymmetric parameter is highest for small refractive index and systematically decreases with increasing n (Fig. 36). Physically this may be explained by the strong forward scattering at small refractive index and the increase of the wave bending away from forward direction at high n. When n , and large , cos 0.5, because half of the scattered radiation is diffracted in the forward direction and half is reflected isotropically (Hansen and Travis, 1974). The Figures 34 and 36 clearly show that for the size parameter > 3, the bulk of scattered light is contained in the forward scattering lobe with < 45 (cos > 0.7). 61 Figure 35. Asymmetry parameter, <cos >, as a function of the effective size parameter, 2a/. The standard size distribution (2.56) was used with four values of the effective variance b. For the case b = 0, 2a/ = 2r/ x. The refractive index is nr = 1.33, ni = 0. 62 Figure 36. Asymmetry parameter, <cos >, as a function of effective size parameter, 2a/. Results are shown for five values fothe real refractive index, nr, all with ni = 0. The standard size distribution (2.56) was used with b = 0.07. Phase Function The directional dependence of the light scattered by a particle is determined by the phase function. For homogenous spheres (uniform refractive index inside the particle) the phase function is determined by the optical size parameter, , the refractive index m = n-in-1, and the scattering angle . For isotropic homogeneous spheres the scattering S ( ) 0 S 1 0 S2 ( ) matrix has the simple form (2.34) 63 It follows that the Stokes parameters, Equation (1.4) of the incident and scattered radiation are related by I= 1/ (k2R2) FI0 (2.35) where, following van de Hulst (1957), we use the symbol F for the four-by-four transformation matrix 1 (S 1 S 1* 12 (S S * F 2 1 1 S 2 S *2 ) (S 1 S 1* S 2 S *2 ) 1 0 0 S 2 S *2 ) (S 1 S 1* S 2 S *2 ) 2 1 i * (S 1 S 2 S 2 S 1* ) (S 1 S *2 0 0 2 2 i 1 (S 1 S *2 S 2 S 1* ) (S 1 S *2 0 0 2 2 2 1 0 0 * S 2 S1 ) * S 2 S 1 ) (2.36) The transformation matrix is proportional to the phase matrix, F = cP (2.37) The proportionality constant follows from the normalization condition on P, Equation (2.5), which yields c F11 4 d 4 (2.38) and the definition of the scattering cross-section sca IR 2 d Io 4 1 F11d 2 k 4 (2.39) Thus C = (k2sca) / 4 (2.40) and the specific relations between the matrix elements are F11 = ((k2sca) / 4) p11 = (1/2) (S1S1* + S2S2*), F21 = ((k2sca) / 4) p21 = (1/2) (S1S1* - S2S2*), F33 = ((k2sca) / 4) p33 = (1/2) (S1S2* + S2S1*), F43 = ((k2sca) / 4) p43 = (i/2) (S1S2* - S2S1*), 64 S1* -- conjugate complex of S1 S2* -- conjugate complex of S2 (2.41) The scattering functions S1 () and S2 ( ) were defined previously (p. 71). P gives the angular distribution and polarization of scattered light for any polarization of incident radiation. The element in the first row and first column, p11 is the phase function and gives the probability of scattering of unpolarized incident light. The ratio p21/p11 is the linear polarization for unpolarized incident light. The significance of the other two matrix elements is indicated in Table 4. For unpolarized incident light p33 and p43 vanish, i.e. only linear polarization occurs. As noted earlier in such a case the intensity functions i1 and i2 completely describe the angular distribution of scattering. The intensity functions i1 an i2 are shown in Fig. 25 for m = 1.33, 1.55, and 2.0, and for the size parameters x = 1, 1.5, 2, 2.5, 3, 2.6, 4, 5, and 6 (van de Hulst, 1957, p. 152). The intensity functions i1 and i2 are shown in a polar diagram for = 1 and for = 10 (Fig. 26). The dominant lobe is confined to < 45, and it is also called the diffraction lobe. With increasing other, smaller lobes occur such that the total number of lobes is approximately . Polarization For small size parameters, << 1, Rayleigh scattering occurs, thus there is a strong positive linear polarization (i1 – i2) / (i1 + i2) > 1 with the maximum polarization at scattering angle = 90o. With increasing the degree of polarization generally decreases and exhibits fluctuations with the major peak shifting towards larger . For ~ 1 m for instance, the peak of the polarization is at ~ 150o (Fig. 38). Since polarization is a ratio measurement, it can be determined accurately. As pointed out by Coulson (1974), it has not been explored adequately in the measurement of atmospheric aerosol parameters. Some aspects of aerosol polarization are discussed by Harris (1972), Ward et al (1973), and White (1975). Aureole Scattering For large size parameters the phase function approaches that for geometrical optics. At small , the phase function is large and the linear polarization vanishes (i1 = i2), because of the predominance of the unpolarized diffracted light. The scattered light contained in the first (forward scattering) lobe is referred to as aureole scattering. Aureole scattering by aerosols is the cause of circumsolar radiation and plays a major role in atmospheric optics. As seen in Fig. 26, the aureole scattering becomes narrower as the particle size increases. For that reason, the aureole scattering function with the sun as the light source, may be used as a measure of the effective aerosol size and the size distribution (Gorchakov et al 1970; Green et al, 1971; Ward et al. 1973; Angstrom, 1974a; Angstrom, 1974b; Gorchakov & Isakov, 1974; Twitty et al. 1976). The circumsolar light scattering may also introduce errors in the turbidity measurements as pointed out by Grassel (1971). 65 Fgure 37. A sample of the scattering diagrams computed by means of the rigorous formulae; m = refractive index, x = 2a/. In all graphs the logarithms of the intensities (1 division = a factor 10) are plotted against the scattering angle (1 division = a factor 10) are plotted against the scattering angle (1 division = 30. Spheres (formulae in sec.9.31): Solid curves i1, dotted curves i2. The values for = 0 and 180 are indicated in the margin. For m = 2, 1.55, and 1.33 they have been obtained by squaring the moduli tabulated by Rayleigh (Sci. Papers 344, 1910). (Yan de Hulst, 1957) 66 Figure 38. Phase fucntion, P11, and percent polarization. -100 P21/ P11, for single scattering of unpolarized incident light. Results are shown for the four size distributions illustrated in the inset, all of which have the same value for reff (1 ) and veff (0.25), where reff is the effective radius and veff the effectie variance. The calcualtions are for the real refractive index nr = 1.33 and wavelength L. (Hanson and Travis, 1974). 67 Figure 39. Model size distriibutions for a Junge (v = 3); an equivalent distribution of four log-Gaussian components; and a Diermendjian Haze. (Harris and McCormics 1972). 68 Figure 40. a.) Average scattering function per particle as a fucntion of angle for parallel polarization and individual log-Gaussian components numbered from 1 (small) to 4 (large) for Mainz aerosol, m = 1.500 - 0.010i at = 0.53 m. b.) Average scattering function per particle as a function of angle for perpendicular polarization and individual log-Gaussian comoponents numbered from 1 (small) to 4 (large) for Mainz aerosols, m = 1.500 - 0.010i at = 0.53 m. Glory Scattering The enhanced intensity in the backscattering direction ~ 180o, is the so-called “Glory.” This is caused specifically by the spherical shape of the scatterers which serves to focus certain rays at ~ 180o. There are essentially two origins for these rays: edge or grazing rays which set up surface waves on the sphere, and noncentral rays which emerge at ~ 180o after internal reflection. The surface waves are not included in the formulation of geometrical optics, and their contribution decreases as the particle size increases. For refractive indices in the range 2 < n < 2, a noncentral ray can emerge at = 180o after just one internal reflection; this gives rise to the intense glory. Bryant & Jarmie (1974) give a good detailed discussion of the glory. Backscattering has received considerable interest recently because of its unique role in remote probing of atmospheric aerosol vertical structure by lidar. An interesting feature of the glory is that it is absent for nonspherical particles and thus it can be sued to distinguish spherical and nonspherical particles. Other features of the phase function are discussed by van de Hulst (1959) and Hansen & Travis (1974). In atmospheric optics we always have to consider the role of aerosol distribution. Numerical work of Hansen & Travis (1974) showed that the effective phase function and the percent linear polarization is similar for aerosol size distributions described by gamma function, log-normal, bimodal, and power law distributions provided that their effective radius (reff) and effective variance (veff) is the same. Their sensitivity analysis was performed for reff = 1 m and veff = 0.25 (Fig. 27). 69 The effect of the size distribution and percent polarization was also studied by Harris & McCormick (1972). They compared the power law distribution, Junge (1963), with = 4, Deirmendjian (1969), haze L distribution, and four log-normal distributions with logmean radius of 0.08, 0.24, 0.64, and 2.0 m (Fig. 28). The phase functions arising from the four log-normal distributions (labeled 1 to 4 with increasing size) are shown for the parallel and the perpendicular polarization components in Fig. 29. The smallest size components (No. 1) approaches Rayleigh scattering, while the largest size (No. 4) clearly illustrates the dominance of the forward scattering lobe. It is instructive to compare the log-normal distributions 1-4 to the two modes of the bimodal distribution (Fig. 10, p. 29). The distribution No. 1 with log-mean radius rgo= 0.08 m for the distribution function corresponds to rg3 = 0.15 m for the volume distribution function. This compares well with the measured volume mean diameter Dp=0.3 m for the fine particle fraction of the Los Angeles smog (Fig. 13, p. 31) and Dp=0.45 for the Nagoya aerosol (fine particle mode). Hence, the phase function for size distribution No. 1 by Harris & McCormick is probably representative for the fine particle mode of urban aerosols. For such a distribution the backward scattering intensity = 180o is 1/10 of the forward intensity and the polarization ratio (i1-i2) / (i1+i2) ~ 10 at ~100o. The distribution No. 4 of Harris & McCormick with volume-mean diameter Dp=7.5 m is illustrative for the coarse particle mode (Fig. 10, p. 29). For such a distribution the bulk of the light scattering is confined to the aureole scattering ( < 10o) while the backscattering intensity is three orders of magnitude smaller. Nonspherical Particles The theory of nonspherical particles is only developed for a few well defined shapes, such as ellipsoids and cylinders, illuminated perpendicularly to their axes (van de Hulst, 1954; Kerker, 1969). The theory scattering by irregularly shaped randomly oriented particles is inhibited by the mathematical problems of satisfying the Maxwell equations, the boundary conditions of the individual nonspherical particle. However, recently a set of experiments on such particles was reported which deserves attention. Measurements of angular scattering from nonspherical particles comparable in size to the wavelength have been primarily on polydispersed systems of particles. Powell et al. (1967) showed that such measurements on magnesium oxide cubes and needlelike fourling of zinc oxide could be approximated with Mie theory for some suitable size distributions of spheres of the same refractive index. Holland’s & Gagne’s (1970) measurements of light scattered by platelike particles of silicon dioxide show good agreement with the Mie theory in the first 50o from forward scattering, but the backscattered radiation is generally less than the theory predicts for an equivalent size distribution of spheres. Scattering measurements by Napper & Ottewill (1962) on monodispersed sols of octahedral and cubic silver bromide exhibit good agreement with Mie theory for the octahedral sols, but generally poor agreement for the cubic sols. These measurements were made only for scattering angles 60-120o from the direction of forward scattering. Scattering measurements on monodispersed Escherichia Coli cells (Cross & Letimer, 1972), which are prolate ellipsoids, and measurements on monodispersed particles of barium sulfate (Peters & Dezelic, 1975) show fair agreement 70 with Rayleigh-Debye theory, providing the real refractive index is close to 1. Less definitive measurements of angular scattering by nonspheres have been made by Quiney and Carswell (1972), by Phillips & Wyatt (1972), and by Hodkinson (1962). Pinnick et al. (1976) measured the polarized light scattered from monodispersed nonspherical randomly oriented aerosol particles and compared with Mie calculations for spheres of approximately the same cross sectional area. They find that for slightly nonspherical particles of sodium chloride and potassium sulfate with size parameter greater than about five, the intensity of light scattered is generally more than as predicted by Mie solutions in the forward scattering lobe less than five, the MIE results more closely match the measurements. Measured angular scattering patterns for randomly oriented particles are smoother than the MIE calculation and are nearly the same for NaCl and potassium sulfate particles of the same size. The angular scattering by polydispersed crystals of potassium chloride particles was measured by Chyler et al. (1976). Their main conclusion is that backscattering is significantly reduced for nonspherical particles as shown in Fig. 31. Chyler et al. (1976) has also proposed a somewhat dubious “theory” for nonspherical particles by truncating the Mie series at specified number of terms. This approach appears to be rather weak on physical grounds. The above experiments consistently show a reduction of the backscattering for nonspherical particles. The primary importance of these observations is that the backscattering of the lidar signals in dry, dusty areas needs to be corrected for this effect. Also, the hemispherical backscattering is less for such particles. 71