4 – Air Pollution Meteorology




Abstract




Population exposure to air pollution occurs mostly near the Earth’s surface. Furthermore, most air pollution sources are located near the Earth’s surface (some exceptions include tall stacks, aircraft emissions, and volcanic eruptions). Therefore, the meteorological phenomena of the lower layers of the atmosphere are the most relevant to understand and analyze air pollution. The part of the atmosphere that is in contact with the Earth’s surface and is affected by it is called the atmospheric planetary boundary layer (PBL). This chapter describes the dynamic processes that take place within the PBL. Those include, in particular, turbulent atmospheric flows and heat transfer processes, which affect air pollution near the surface. Those processes are often referred to as “air pollution meteorology,” because they are the most relevant to air pollution. The major equations governing these processes are presented. A more detailed description of the PBL is available in books such as those by Stull (1988) and Arya (2001).





4 Air Pollution Meteorology



Population exposure to air pollution occurs mostly near the Earth’s surface. Furthermore, most air pollution sources are located near the Earth’s surface (some exceptions include tall stacks, aircraft emissions, and volcanic eruptions). Therefore, the meteorological phenomena of the lower layers of the atmosphere are the most relevant to understand and analyze air pollution. The part of the atmosphere that is in contact with the Earth’s surface and is affected by it is called the atmospheric planetary boundary layer (PBL). This chapter describes the dynamic processes that take place within the PBL. Those include, in particular, turbulent atmospheric flows and heat transfer processes, which affect air pollution near the surface. Those processes are often referred to as “air pollution meteorology,” because they are the most relevant to air pollution. The major equations governing these processes are presented. A more detailed description of the PBL is available in books such as those by Stull (1988) and Arya (2001).



4.1 The Atmospheric Planetary Boundary Layer


The PBL can be defined according to its dynamic or thermal processes. From a dynamic viewpoint, the PBL is the atmospheric layer that is affected by the Earth’s surface, either because of mechanical effects due to terrain, buildings, vegetation, etc., or indirectly because of heat transfer processes related to the evaporation of water or the anthropogenic production of heat (for example, the urban heat island effect). From a thermal viewpoint, the PBL is the atmospheric layer that is affected by the temporal variation of solar radiation, for example via its ambient temperature.


The PBL height varies as a function of terrain and solar radiation. It is on the order of a kilometer, but it is lower in winter when there is little solar radiation (on the order of a few hundred meters) and greater in summer when solar radiation is maximum (on the order of 2,000 to 3,000 m).


Within the PBL, one distinguishes the surface layer, which is located near the surface. Within the surface layer, vertical fluxes of momentum, heat, and mass transfer are nearly constant (~±10 %). The surface layer has a thickness that is less than 10 % of that of the PBL. One should note, however, that within a forest or an urban canopy, the atmospheric flows are generally complex and the vertical fluxes are most likely not constant.


Above the PBL, the effect of the Earth’s surface is negligible: this is the free atmosphere. When one is interested in air pollution at local or urban scales, it is generally appropriate to take into account only the PBL. On the other hand, when one is interested in air pollution at regional, continental, or global scales, the free troposphere must be taken into account, because air pollutants can be transported over long distances in the free troposphere and precipitation events, which scavenge air pollutants, generally include a significant fraction of the free troposphere.



4.2 Vertical Motions of Air Parcels



4.2.1 Potential Temperature


The change of pressure and temperature as a function of altitude is obtained via the hydrostatic and ideal gas laws (see Chapter 3). Furthermore, a relationship between pressure and temperature may be obtained by assuming that air parcels undergo adiabatic changes. Therefore, the law of Laplace applies:



PVγ = constant
PVγ=constant
(4.1)

where γ is the adiabatic index of the gas (here, air), also known as the Laplace coefficient (γ = cp / cv, where cp and cv are the specific heat capacities at constant pressure and volume, respectively). Applying the ideal gas law (see Chapter 3), the following relationship between pressure and temperature is obtained under adiabatic conditions:



PVγ/(1 − γ) = constant
PTγ/(1−γ)=constant
(4.2)

For a diatomic gas (which is the case of the atmosphere, since molecular nitrogen and oxygen account for 99 % of the air), γ = 7/5 = 1.4. Thus:



PT−3.5 = constant
PT−3.5=constant
(4.3)

Therefore:



T P−0.29 = constant
TP−0.29=constant
(4.4)

The potential temperature, θ, is defined as the temperature of an air parcel brought back to sea level at a standard pressure (i.e., P0 = 1 atm) under adiabatic conditions. Therefore:


TP−0.29=θ P0−0.29(4.5)

Thus:


θ=T(P0P)0.29(4.6)


4.2.2 Neutral Atmosphere


A constant vertical profile of the potential temperature is defined as neutral. Let us consider an air parcel with uniform properties (pressure, temperature, and density). Under neutral conditions, this air parcel will remain in equilibrium with its environment as it moves upward or downward, since its temperature will change adiabatically and will, therefore, remain consistent with the temperature of the surrounding air.



4.2.3 Unstable Atmosphere


The atmosphere is unstable if the vertical profile of the potential temperature is negative, i.e., if the actual temperature decreases faster with height than the adiabatic gradient. An air parcel moving upward becomes warmer than that of its surroundings, because its adiabatic change in temperature is less than the decrease of the surrounding atmospheric temperature. Therefore, according to the ideal gas law, the density of this air parcel becomes less than that of the surrounding air. Thus, this air parcel keeps moving upward. Similarly, an air parcel moving downward undergoes an adiabatic change and its temperature change is less than that of the gradient of the surrounding atmosphere. Therefore, the surrounding temperature becomes greater than that of the air parcel. Thus, this air parcel is denser than the surrounding air and it continues to move downward. Vertical motions in an unstable atmosphere are enhanced because a small perturbation in the vertical direction leads to continuous motion in the direction of the initial perturbation (see Figure 4.1). Such unstable air parcels lead to vertical mixing of air pollution, at least up to an altitude where the atmosphere becomes neutral or stable (see Section 4.2.4). An unstable atmospheric layer is called a mixing layer. It occurs generally within the PBL during the day, because solar radiation warms up the Earth’s surface, but does not affect the temperature of the lower atmosphere (since the air absorbs little ultraviolet radiation within the troposphere, see Chapter 5). As a result, the vertical potential temperature gradient becomes negative.





Figure 4.1. Schematic representation of vertical motions of an air parcel for neutral conditions (left figure), unstable conditions (middle figure), and stable conditions (right figure). The solid arrows indicate the initial perturbation of the air parcel, which undergoes an adiabatic change, and the dashed arrows indicate the following motion of the air parcel, which is due to the difference between the temperature of the air parcel and that of the surrounding air (i.e., density difference).



4.2.4 Stable Atmosphere


The atmosphere is stable if the vertical profile of the potential temperature becomes positive. As an air parcel moves upward, its temperature follows an adiabatic change and it becomes lower than that of the surrounding air. Therefore, the air parcel becomes denser than the surrounding air. As a result, it tends to move back downward. Similarly, an air parcel moving downward undergoes an adiabatic change and its temperature becomes greater than that of the surrounding air. Thus, this air parcel, being less dense than the surrounding air, tends to move back upward. Therefore, a stable atmosphere tends to suppress vertical motions of air parcels. Thus, such an atmospheric layer may become stratified, because there are no, or few, transfers between the different layers (strata) of the atmosphere (see Figure 4.1). A stable atmospheric layer may be present aloft, for example above a mixing layer, or at the surface. This latter case occurs at night when the infrared radiation emitted by the Earth’s surface toward the atmosphere leads to a decrease of the surface temperature and, accordingly, of the surface layer temperature. Therefore, a temperature inversion develops: the gradient of the actual temperature becomes positive and the gradient of the potential temperature becomes also positive. (Note that an atmospheric layer may be stable without a temperature inversion, as long as the actual temperature gradient is less than the adiabatic gradient.) Then, the atmospheric layer near the surface is stable at night. This stable layer will gradually become thicker as the surface temperature keeps decreasing. The PBL layer located above this stable layer is called the residual layer. Pollutants located within this residual layer are then isolated from the surface. However, during daytime, as solar radiation warms the surface, thereby leading to the gradual disappearance of the surface-based stable layer, the previously stable and residual layers merge to form a new mixing layer. Pollutants that were formerly present within the residual layer are mixed vertically and may then interact with the surface. The evolution of the atmospheric layer near the Earth’s surface is depicted schematically in Figure 4.2. The stratosphere shows a positive vertical profile of the potential temperature; therefore, it is a stable atmospheric layer with few vertical air motions.





Figure 4.2. Schematic representation of the daily cycle of the planetary boundary layer.


Source: After Stull (1988).


4.2.5 Effect of Moisture


This description of temperature gradients applies to dry atmospheres. If the atmosphere is moist, its vertical temperature gradient under adiabatic conditions will be less than that of a dry atmosphere. The water content, q (i.e., the specific humidity expressed in g of water per g of moist air), corresponds to a relative humidity (%), which is greater when temperature decreases, because the saturation vapor pressure of water decreases with temperature. In other words, for a given quantity (mass) of water per quantity of air (mass or volume), relative humidity increases when temperature decreases. Accordingly, for a given relative humidity, the water content of the atmosphere decreases when temperature decreases.



Humidity

Relative humidity (%) may be converted to absolute humidity using the method proposed by McRae (1980). According to Dalton’s law, the water vapor concentration expressed in parts per million (ppm; 1 ppm = 1 molecule per million molecules of air) is:


[H2O(ppm)]=104RH(%)Ps,H2O(T)P(4.7)

where T is the ambient temperature, Ps,H2O(T) is the saturation vapor pressure of water, and P is the atmospheric pressure. The saturation vapor pressure of water depends on temperature according to the following polynomial formula:


Ps,H2O(T)=P exp(13.3185ωT−1.9760ωT2−0.6445ωT3−0.1299ωT4)(4.8)

where ωT = 1 – (373.15/T), where T is expressed in K. For example, at P = 1 atm and T = 25 °C, Ps,H2O (298 K) = 0.031 atm; at T = 5 °C, Ps,H2O (278 K) = 0.0086 atm. At a relative humidity of 50 % and a temperature of 5 °C, there are about 4300 ppm of water vapor.


Converting % to (g m−3) is performed as follows:


[H2O(g m−3)]=0.18 RH(%)Ps,H2O(T)R T(4.9)

Thus, a relative humidity of 50 % at 1 atm and 5 °C corresponds to 3.4 g of water vapor per m3 of air. The water vapor mass per volume of air is called the volumetric absolute humidity. The water vapor mass concentration at saturation (RH = 100 %) increases with temperature. For example, at 1 atm, there are 9.4, 17, and 30 g of water vapor per m3 at 10, 20, and 30 °C, respectively.


Absolute humidity may also be defined as the mass of water vapor per mass of air. It can be defined per mass of dry air or per mass of moist air. The absolute humidity per mass of dry air is usually called the mixing ratio (by weight) and the absolute humidity per mass of moist air is usually called the specific humidity. The molar mass of dry air is 29 g mole−1; therefore, according to the ideal gas law, the mass in g of one m3 of dry air is: 29 Pdry air /(RT). The mixing ratio is thus related to relative humidity as follows:


[H2O(g/g dry air)]=0.18RH(%)Ps,H2O(T)R TR T29 Pdry air=0.62RH(%)100Ps,H2O(T)Pdry air(4.10)

According to Dalton’s law, the total pressure is the sum of the partial pressures of dry air and water vapor:


P=Pdry air+RH(%)100Ps,H2O(T)(4.11)

Therefore:


[H2O (g/g dry air)]=0.62RH(%)100Ps,H2O(T)(P−RH(%)100Ps,H2O(T))(4.12)

The specific humidity, q, is related to relative humidity as follows:


q=1(1.61PPs,H2O(T)100RH(%))−0.61(4.13)

Given that water vapor is a small fraction of the total mass of an air parcel, the mixing ratio and the specific humidity generally have very close values. For example, at T = 25 °C and P = 1 atm, Ps,H2O = 0.031 atm (see previous equation). For a relative humidity of 100 %, the mixing ratio is calculated to be 0.0198 g H2O/g dry air and the specific humidity, q, is 0.0195 g H2O/g moist air, i.e., a difference of less than 2 %.



Virtual Temperature

The virtual temperature of a moist air parcel, Tv, is defined as the temperature that dry air would have for the same values of pressure and density as moist air. The relationship between the actual temperature, T, and the virtual temperature, Tv, (in K) is a function of the specific humidity, q (g of water/g of moist air), of the air parcel. It is given approximately by the following formula (see, for example, Wallace and Hobbs, 2006, for the exact relationship and the derivation of this simplified relationship):



Tv = T(1 + 0.61q)
Tv=T(1+0.61q)
(4.14)

The ideal gas law applies to a moist air parcel if Tv is used instead of T. Since water vapor has a molar mass (18 g mole−1) that is less than that of dry air (29 g mole−1), unsaturated moist air is less dense than dry air. Therefore, according to the ideal gas law, the virtual temperature is always greater than the actual temperature (by a few degrees at most), as shown by Equation 4.14. A virtual potential temperature, θv, may be defined as the potential temperature that dry air would have for the same values of pressure and density:



θv = θ(1 + 0.61q)
θv=θ (1+0.61q)
(4.15)


Vertical Temperature Gradients

The adiabatic gradient of saturated moist air is, in absolute value, less than that of dry air. This gradient is due to the fact that at saturation, a moist air parcel moving upward will cool down and, therefore, will undergo partial condensation of water. Since condensation produces heat, the temperature of the air parcel decreases less than in the case of dry air. The atmospheric conditions can be categorized as follows in terms of the vertical temperature gradients for dry air and air saturated with water (the gradients are expressed here in absolute values and for cases where temperature decreases with altitude):




  1. Temperature gradient > dry adiabatic gradient: unstable atmosphere



  2. Temperature gradient = dry adiabatic gradient: neutral atmosphere if dry



  3. Saturated adiabatic gradient < temperature gradient < dry adiabatic gradient: stable, neutral, or unstable atmosphere depending on humidity



  4. Temperature gradient < saturated adiabatic gradient: stable atmosphere



4.3 Winds and Turbulence



4.3.1 Wind Direction in the PBL


As discussed in Chapter 3, the wind direction in the free troposphere results from the balance between two forces: the pressure gradient force and the Coriolis force (which is due to the Earth’s rotation). In the PBL, another force comes into play: the frictional force, which represents the effect of the friction of the flow on the Earth’s surface. Therefore, the wind direction at the surface results from the balance among these three forces: the pressure gradient force, the Coriolis force, and the frictional force. The frictional force is opposite to the wind direction. Furthermore, the Coriolis force is proportional to the wind speed. Since the wind speed decreases near the surface (tending toward zero at or near the surface, see Section 4.3.2), the Coriolis force also decreases near the surface. Thus, the influence of the pressure gradient force on the wind direction is greater near the surface. The theoretical solution for the northern hemisphere leads to a wind direction near the surface that is 45 ° to the left of the geostrophic wind direction. Thus, the wind direction evolves from that direction near the surface toward the geostrophic wind direction at the height within the PBL where the frictional force becomes negligible. The change in wind direction as a function of height is called the Ekman spiral. The depth of this Ekman layer, where the change in wind direction occurs, depends on the effect of the surface. It is greater for a situation with complex terrain (urban canopy, forest …) than for a smooth surface (sea, snow, ice …). The equations that govern the vertical profile of the wind speed and direction are described next.



4.3.2 Equations of Atmospheric Flow and Turbulence


The Navier-Stokes equations are used to calculate the atmospheric flow. They are expressed in Cartesian coordinates as follows:


∂u∂t+u∂u∂x+v∂u∂y+w∂u∂z−f v=−1ρa∂P∂x+1ρa (∂τxx∂x+∂τyx∂y+∂τzx∂z)∂v∂t+u∂v∂x+v∂v∂y+w∂v∂z+f u=−1ρa∂P∂y+1ρa (∂τxy∂x+∂τyy∂y+∂τzy∂z)∂w∂t+u∂w∂x+v∂w∂y+w∂w∂z      =−1ρa∂P∂z+ 1ρa (∂τxz∂x+∂τyz∂y+∂τzz∂z)−g(4.16)

where the terms τij represent the shear stresses (kg m−1 s−2), which are due to the fluid viscosity. For a newtonian fluid, Newton’s law of viscosity applies. It is written for fluid motion in a single direction as follows:


τyx=μv,a∂u∂y(4.17)

where μv,a is the dynamic viscosity of the air (kg m−1 s−1). This equation may be seen as the flux in the y direction of the momentum in the x direction. This law, defined for a fluid moving in a given direction (here, x), has been generalized by Navier, Stokes, and Poisson for a fluid moving in three dimensions. Then, the shear stresses for momentum in the x direction are as follows:


τxx=2μv,a ∂u∂x+(μD−23μv,a) (∂u∂x+∂v∂y+∂w∂z)τyx=μv,a (∂u∂y+∂v∂x)τzx=μv,a (∂u∂z+∂w∂x)(4.18)

The other terms for the y and z directions, τxy, τyy, τzy, τxz, τyz, and τzz, are written similarly. Clearly: τxy = τyx, τyz = τzy, and τxz = τzx. In these equations, μv,a is the dynamic viscosity of the fluid and μD is the dilatational viscosity of the fluid (μD = 0 for a monoatomic gas and μD ≈ 0 for air).


Boussinesq introduced several assumptions that apply to the PBL:




  1. The kinematic viscosity and the molecular thermal conductivity are constant (i.e., their dependence on pressure and temperature can be neglected).



  2. The heat produced by the viscous forces is negligible compared to solar radiation in the heat transfer equation.



  3. The atmosphere is considered to be incompressible, since it is an open system (except at the Earth’s surface).



  4. The fluctuations of the state variables (pressure, temperature, and density) are small compared to their mean values.



  5. The ratio of the pressure fluctuation and the mean pressure is small compared to the ratio of the temperature fluctuation and the mean temperature and compared to the ratio of the density fluctuation and the mean density.



  6. The fluctuations of the density are important only when the density is multiplied by the gravitational constant; indeed, the Earth’s surface introduces a constraint on the system in the vertical direction.


The last two assumptions are commonly called the Boussinesq approximation or hypothesis: they imply that the change in the air density can be neglected except in the term (ρa g dz) of the hydrostatic equation (see Equation 3.2).


Using the Boussinesq hypothesis, the continuity equation for momentum is as follows:


∂u∂x+∂v∂y+∂w∂z=0(4.19)

For u, the friction term may be written as follows (assuming that μv,a is spatially uniform):


1ρa(∂τxx∂x+∂τyx∂y+∂τzx∂z)=μv,aρa(∂∂x(2∂u∂x−23(∂u∂x+∂v∂y+∂w∂z))+∂∂y (∂u∂y+∂v∂x)+∂∂z (∂u∂z+∂w∂x))(4.20)

Changing the order of the derivations:


1ρa(∂τxx∂x+∂τyx∂y+∂τzx∂z)=μv,aρa((∂2u∂x2+∂2u∂y2+∂2u∂z2)−23∂∂x(∂u∂x+∂v∂y+∂w∂z)+∂∂x(∂u∂x+∂v∂y+∂w∂z))(4.21)

According to the continuity equation, the last two terms are zero. Since the kinematic viscosity, νv,a(m2 s−1), is equal to the ratio of the dynamic viscosity and the fluid density, μv,a/ρa:


1ρa(∂τxx∂x+∂τyx∂y+∂τzx∂z)=νv,a (∂2u∂x2+∂2u∂y2+∂2u∂z2)(4.22)

Thus, the Navier-Stokes equations may be written as follows:


∂u∂t+u∂u∂x+v∂u∂y+w∂u∂z−fv=−1ρa∂P∂x+νv,a ∇2u∂v∂t+u∂v∂x+v∂v∂y+w∂v∂z+fu=−1ρa∂P∂y+νv,a ∇2v∂w∂t+u∂w∂x+v∂w∂y+w∂w∂z   =−1ρa∂P∂z+νv,a ∇2w−g(4.23)

where u, v, and w are the wind velocities (m s−1) in the x, y, and z directions, respectively, f is the Coriolis parameter (s−1), ρa is the air density (kg m−3), P is the atmospheric pressure (Pa), νv,a is the kinematic viscosity of the air (m2 s−1), and g is the acceleration due to gravity (m s−2).


The Reynolds number, Re, may be used to estimate whether a flow is turbulent or laminar. It represents the ratio of inertial forces and viscous forces, and it is defined as follows:


Re=ulcνv,a(4.24)

Only gold members can continue reading. Log In or Register to continue

Oct 12, 2020 | Posted by in General Engineering | Comments Off on 4 – Air Pollution Meteorology
Premium Wordpress Themes by UFO Themes