Open access peer-reviewed chapter

A Novel Development in Three-Dimensional Analytical Solutions for Air Pollution Dispersion Modeling

Written By

Mehdi Farhane and Otmane Souhar

Reviewed: 19 June 2023 Published: 18 July 2023

DOI: 10.5772/intechopen.112225

From the Edited Volume

Air Pollution - Latest Status and Current Developments

Edited by Murat Eyvaz, Ahmed Albahnasawi and Motasem Y. D. Alazaiza

Chapter metrics overview

119 Chapter Downloads

View Full Metrics

Abstract

A novel analytical solution has been developed for the three-dimensional dispersion of atmospheric pollutants, providing a new approach to understanding and addressing this important environmental issue. The central concept of the study is to divide the planetary boundary layer into multiple vertical sub-layers, each characterized by its own average wind speed and eddy diffusivity. This allows for a more comprehensive and nuanced examination of atmospheric processes within the boundary layer. The validity of the model is thoroughly evaluated through a comparison of its predictions with data collected from the Copenhagen Diffusion and Prairie Grass experiments. This approach ensures that the model accurately reflects the complexities of atmospheric dispersion in real-world scenarios. The results of the study demonstrate a strong correlation between the predicted and measured crosswind-integrated concentrations. Furthermore, the statistical indices computed for the model fall within an acceptable range, indicating a high level of accuracy in the model’s predictions. These findings reinforce the validity of the analytical solution for modeling atmospheric pollutant dispersion.

Keywords

  • analytical solution
  • atmospheric dispersion
  • atmospheric boundary layer sub-layers
  • Fourier transform
  • Sturm-Liouville eigenvalue problem

1. Introduction

The scientific community has been driven to act urgently due to the increase in ecological disasters worldwide. A key approach to addressing this issue is the development of reliable models that enable quantitative prediction of pollution-related phenomena, either through analytical descriptions or simulations using powerful and operational tools. These models, which can be based on simulations or analytical descriptions, have strong quantitative predictive power for understanding pollution-related phenomena. The atmosphere is considered the primary means of dispersing pollutants in the environment, which can come from industrial sources or accidental events, leading to progressive contamination of the ecosystem, fauna, flora, and all populations. Therefore, accurately evaluating how pollutants move through the atmosphere in the boundary layer is vital to preserve, protect, and restore the integrity of ecosystems. To accomplish this, it is essential to create atmospheric dispersion models tailored to the parameters and weather conditions specific to the region being studied, using parameters, weather conditions, and local topographic information. These models should produce realistic outcomes on environmental consequences, helping to minimize the negative impact of potential disasters like forest fires. Furthermore, implementing simulations of models based on specific cases can be beneficial in establishing particular emission limits for industrial sites, limiting the release of pollutants into the air.

The aim of our work is to propose a closed-form analytical solution for three-dimensional advection-dispersion transport problems in finite, multilayered media using rigorous mathematical tools. We ensure that wind velocity profiles and vertical diffusivity coefficients take average values in each sub-layer. We have made innovative contributions by using a governing function to generate the eigenvalues associated with our problem and overcoming the common difficulty of missing some of the eigenvalues when they are calculated by developing a transcendental equation for each layer. Finally, our approach has enabled us to obtain a closed-form analytical solution that differs from previous solutions that required the determination of integral coefficients. This method could have important implications in many scientific and technical fields for solving complex advection-dispersion transport problems in finite, multilayered media.

In recent years, there has been an increasing emphasis on creating new analytical approaches that can be used for a range of wind speeds and turbulent diffusivity coefficients. For the most part, these approaches involve the projection of the solution onto a basis of orthogonal polynomials, such as the GILTT technique [1, 2, 3, 4, 5, 6, 7, 8]. However, an important drawback of this method is that it requires a large number of eigenvalues to ensure convergence, which can reach up to 250 eigenvalues. In contrast, our new solution is able to provide better results by using a much smaller number of eigenvalues, specifically between 10 and 15, to ensure convergence. This is a significant improvement over existing approaches, making our method more efficient and practical for analytical applications.

Our model was developed using the Fourier transform and separation of variables technique, which resulted in the Sturm-Liouville problem. In order to understand the factors influencing pollutant dispersion, we considered the following parameters: (i) the wind speed profile of Deaves and Harris [9], (ii) the vertical turbulent diffusivity coefficient, which is considered as an explicit function of the downwind distance and vertical height under convective conditions, as described by Mooney and Wilson [10] and Degrazia et al. [11], and (iii) the lateral eddy diffusivity coefficient, which also depends on the downwind distance and vertical height, as described in the work of Huang [12] and Brown et al. [13]. We first proceed to the exposition of the general formulation of the equation that governs the dispersion of pollutants in the atmospheric boundary layer, with a comprehensive presentation in Section 2. We then turn to the explicit solution in Section 3, while exploring the model parameterizations in Section 4. We provide an in-depth discussion of our numerical results in Section 5. Finally, the last section proposes our conclusion, a synthesis of our investigation.

Advertisement

2. Presentation of the general problem

The movement of pollutants within the planetary boundary layer (PBL) is characterized by turbulent dispersion, which is governed by a mathematical equation known as the advection-diffusion equation. This equation considers the combined influence of two primary mechanisms, advection and diffusion, which respectively refer to the transport of pollutants by wind and the spreading of pollutants due to turbulence.

By applying the advection-diffusion equation, researchers can gain a more comprehensive understanding of the intricate and dynamic behavior of pollutants within the PBL. This equation serves as a mathematical framework that accounts for various factors that affect pollutant dispersion, such as wind direction and speed, Eddy diffusivities, and atmospheric stability, and pollutant concentration, presented in the following form:

Ct+.UwC=.DC+S,E1

where Uw=UVWT is the wind speed vector (m/s) representing the components U, V, and W in the east-west, north-south and vertical directions, respectively; D is the molecular diffusion coefficient; S is the source term; and is the gradient operator.

By use of the time average and fluctuation values, U=u+u, V=v+v, W=w+w and C=c+c, the wind speed vector Uw is expressed as:

Uw=Uw¯+Uw',withUw¯=uvwTandUw'=uvwTE2

Applying the Reynolds averaging rules [14] to the vertical mass flow term, denoted as UwC, results in the following expression: It turns out that turbulent diffusion can be described with Fick’s laws of diffusion as follows [15].

uc¯=Kxcx;vc¯=Kycy;wc¯=Kzcz.E3

where Kx, Ky, and Kz are the eddy diffusivities components along x, y, and z directions, respectively.

It should be noted that in a turbulent boundary layer where advection is occurring, K will be larger than D and eddy diffusion will dominate solute transport. In this case, the molecular diffusion coefficient .DC is then to be replaced by an eddy or turbulent diffusivity. The source term could be eliminated from Eq. (1) and should be added to the boundary conditions as a delta function: At the point 00Hs, there is a source rejecting the pollutant with a continuous flow Q:

uc0yz=QδyδzHs,E4

where is the tensor product of two distributions and Hs is the source height. By application of the Reynolds averaging and the divergence operator to Eq. (3), Eq. (1) may be written as follows.

ct=xuc¯yvc¯zwc¯Uw¯.c.E5

In the remainder of this paper, the following assumptions are considered:

  1. the steady state condition (i.e., ct=0);

  2. the two terms vcy and wcz are neglected since the x-axis coincides with the wind flow average, therefore the w and v wind velocity components are less important; and

  3. the turbulent diffusion in the direction of the mean wind is neglected compared to the advection transport mechanism (i.e., ucx>>xKxcx).

These assumptions lead to the steady-state advection-diffusion equation defined as:

uzcx=yKycy+zKzcz;xyz]0,Lx×Ly,Ly×0,Hmix[,subject to the boundary conditions:uzcxyz=QδyδzHs;x=0,Kyxzcxyzy0;yLy,Kzxzcxyzz=0;z0Hmix.E6

where z0 is the surface roughness length and Hmix is the PBL height.

We consider that the eddy diffusivities have the following separable formulations:

Kyxz=ζyxuz,E7
Kzxz=ξxφzz.E8

We vertically divide the PBL into H intervals such that for each one the eddy diffusivity and wind speed assume average values. For h=1,,H,

uh=1zhzh1zh1zhusds,E9
φzh=1zhzh1zh1zhφzsds.E10

By use of the formulations of Ky and Kz given by Eqs. (7) and (8), Eq. (5) is written as:

uhchx=ζyxuh2chy2+ξxφzh2chz2,E11

with uh and φzh (given by Eqs. (9) and (10)) are constants.

This later Eq. (11) is subject to the first boundary conditions Eq. (6) on the one hand, and on the other hand, the continuity of both the concentration and the flux at the interface level is applied.

φz1c1xyz0z=0,ch1xyzh1=chxyzh1,h2H,φzh1ch1xyzh1z=φzhchxyzh1zφzHcHxyzHz=0
Advertisement

3. A mathematical approach to solving the pollutant dispersion equation: analytical solution

We start this section by applying Fourier transform to Eq. (11). Let ĉhωxz denote the Fourier transformation of ch with respect to y.

ĉhωxz=+chxyze2iπωydy,h1H

which gives

uhĉhωx+2π2ω2ζyxĉhω=ξxφzh2ĉhωz2,zzh1zh.E12

Let

χhωxz=ĉhωxz.exp2π2ω20xζysdsE13

then

χhωx=ĉhωx+2π2ω2ζyxĉhωexp2π2ω20xζysds.

By multiplying both sides of Eq. (12) by exp2π2ω20xζysds and since uh and φzh are constants for each interval, we show easily that, for all h1H:

uhχhωx=ξxφzh2χhωz2,zzh1zhE14

We proceed in the same way with the boundary conditions, and find

φz1χ1ωxz0z=0χh1ωxzh1=χhωxzh1φzh1χh1ωxzh1z=φzhχhωxzh1z,h2HφzHχHωxzHz=0

The solution of Eq. (14) is assumed to be in the form:

χhωxz=n=0Gh,nωxPh,nz,h1H

This separated form gives two ordinary differential equations to be solved:

dGh,nωdx+γn2ξxGh,nω=0E15

and

φzhd2Ph,ndz2+γn2uhPh,n=0,E16

where γn is a separation constant.

The first-order ordinary differential Eq. (15) has the solution

Gh,nωx=μnω.expγn20xξsds,

where μn is an arbitrary function depending on ω.

Eq. (16) represents a Sturm-Liouville problem. Solutions of such problem form an eigenfunction basis of the form:

Ph,nz=αh,ncosλh,nz+βh,nsinλh,nz,E17

where, λh,n=γnuhφzh.

The Eq. (17) satisfies the following boundary conditions:

φz1dP1,nz0dz=0aPh1,nzh1=Ph,nzh1φzh1dPh1,nzh1dz=φzhdPh,nzh1dz,h2HbφzHdPH,nzHdz=0cE18

To calculate the expression of Ph,n, it comes down to calculate the values of αh,n and βh,n, on each of the sub-layer zh1zh,h1H.

By solving the recursive system resulting from substitution of Eq. (17) in Eq. (18), we obtain respectively the formulations of αh,n and βh,n. More specifically,

For the first sub-layer, α1,n and β1,n satisfy the equation:

α1,nsinλ1,nz0β1,ncosλ1,nz0=0E19

from which, we can take β1,n=sinλ1,nz0, so that α1,n=cosλ1,nz0, which means:

P1,nz=cosλ1,nzz0.

For the last sub-layer (Hth sub-layer):

αH,n=cotλH,nzHβH,n.E20

And for the intermediate sub-layers:

αh,n=φzhλh,nφzh1λh1,n2φzhλh,ncosλh,n+λh1,nzh1+φzhλh,n+φzh1λh1,nφzhλh,nφzh1λh1,n.cosλh,nλh1,nzh1)αh1,n+(sinλh,n+λh1,nzh1φzhλh,n+φzh1λh1,nφzhλh,nφzh1λh1,nsinλh,nλh1,nzh1βh1,n,E21

and

βh,n=φzhλh,nφzh1λh1,n2φzhλh,nsinλh,n+λh1,nzh1+φzhλh,n+φzh1λh1,nφzhλh,nφzh1λh1,n.sinλh,nλh1,nzh1)αh1,n(cosλh,n+λh1,nzh1φzhλh,n+φzh1λh1,nφzhλh,nφzh1λh1,ncosλh,nλh1,nzh1βh1,n.E22

The eigenvalues γn,n of this problem are real and discrete and the eigenfunctions are mutually orthogonal. The orthogonality relation developed by [16] for this class of (self-adjoint) problems with respect to the density uh on each interval zh1zh,h1H leads to

h=1Hzh1zhuhPh,msPh,nsds=Ph,m.Ph,n.δm,n,

where δm,n is the Kronecker symbol. Then, Ph,n2 is written as follows.

PH,n2=h=1Hzh1zhuhPh,ns2ds={h=1Huhzhzh1,whenγn=0h=1Huh2λh,n[sinλh,nzhzh1(αh,n2βh,n2cosλh,nzh+zh1+2αh,nβh,nsinλh,nzh+zh1)+λh,nαh,n2+βh,n2zhzh1],whenγn0.

The eigenvalues of each sub-layer can be obtained by integrating Eq. (17) on each of the intervals zh1zh,h1H taking into account the boundary conditions Eqs. (19)(21). Whereas the eigenfunctions PH,n form a complete set, so χHω can be developed as:

χHωxz=n=0anexpγn20xξsdsPH,nz.

Then, the coefficients an are given by:

an=1Ph,n2=1Hz1zuχω0zP,nzdz=Qh=1Huhzhzh1,whenγn=0QPh,n2=1HP,nHs,whenγn0

The inverse Fourier transform of exp2π2ω20xζysds is:

12π0xζysdsexpy240xζysds

Consequently,

cHx,yz=Q2π0xζysdsexpy240xζysds.n=0expγn20xξsdsPH,nzPH,n2h=1HPh,nHs.E23

The Crosswind-Integrated Concentration (CIC) model is a valuable tool for analyzing the behavior of pollutants in the atmosphere. It takes into account the horizontal wind component, which is a significant factor in pollutant transport. Mathematically, the crosswind-integrated concentration cHyxz is obtained by integrating cHxyz in Eq. (20) to y from to +:

cHyxz=+cHxyzdy

The (CIC) model considers the dispersion of pollutants in the crosswind direction due to the horizontal wind component and produces a two-dimensional concentration map that can be used to assess the impact of pollutants in both the downwind (x-axis) and crosswind (z-axis) directions by integrating the concentration of pollutants over the crosswind direction (y), it gives

cHyxz=Qn=0expγn20xξsdsPH,nzPH,n2h=1HPh,nHs.E24
Advertisement

4. Parameterizing pollutant dispersion: Exploring model inputs

The wind speed profile u(z) is determined using the Deaves and Harris model [9], which is based on experimental measurements of wind speed profiles. This model is advantageous because it provides a more accurate representation of the logarithmic law at low heights, and improves the accuracy of both the logarithmic and power law models at moderate heights. The profile is given by the equation

uz=uklnz+z0z0+5.75zHmix1.88zHmix21.33zHmix3+0.25zHmix4,E25

where u is the friction velocity, k is the von Karman constant, z0 is the roughness length, and Hmix is the height of the planetary boundary layer (PBL). The Deaves and Harris model includes terms that account for the effects of turbulence and thermal stratification on the wind profile. By using this model, we can better understand and predict the behavior of the wind at different heights within the PBL.

The modified form of the vertical eddy diffusivity coefficient Kz, given in Eq.(8), is adopted from [11], where the functional φzz is expressed as:

φzz=0.22HmixwzHmix1zHmix1/31exp4zHmix0.0003exp8zHmix,E26

where w is the convective velocity. The integrable correction dimensionless function in Eq. (8) is defined in terms of the along-wind length scale L1 as follows [10]:

ξx=1expxL1.E27

The length L1 is given in terms of u, φz and σw as [10]:

L1=1σw2HsuHsφzHs,

where σw is the vertical turbulent intensity. Note that, there exist many expressions of σw, we adopt here the expression given by [17]:

σw=0.96w3zHmix+LHmix1/3,

The Monin-Obukhov length (L) is a commonly parameter used in atmospheric dispersion modeling [18]. It plays a crucial role in characterizing the stability of the atmospheric boundary layer, which is essential for understanding and predicting the dispersion of pollutants or particles in the atmosphere. Specifically, the sign of L indicates the atmospheric stability state. A negative value of L indicates an unstable atmosphere, where turbulence is generated due to buoyancy forces. A positive value of L indicates a stable atmosphere, where turbulence is suppressed due to stratification. A large absolute value of L (i.e., L1) indicates a neutral atmosphere, where turbulence is neither generated nor suppressed due to buoyancy effects [18].

It should be noted that when L10, the term expxL1 becomes negligible in (27), and Kz will dependent only on z (i.e., Kzφz in Eq. (8)).

The lateral eddy diffusivity is given by [12]:

Kyxz=12uzdσy2xdx,E28

where σy is the standard deviation in the crosswind direction (depends only on x).

By identifying Eqs. (7) and (28), we obtain

ζyx=12dσy2xdx
Advertisement

5. Validating analytical solutions for air pollution dispersion modeling: analyzing experimental data

This section offers a comprehensive analysis of the data generated by our model, which is founded on the Deaves and Harris wind velocity profile. The model is constructed using two different formulations of vertical turbulent diffusivity. The first formulation, with L1>0, factors in both the downwind distance and the vertical height, while the second formulation, with L10, considers only the vertical height. The performance of the model, as presented in Eq. (24), is assessed and confirmed through the use of datasets from the Copenhagen diffusion experiments in Denmark and the Prairie Grass experiments in the USA.

Atmospheric dispersion experiments were conducted in Copenhagen, Denmark, from June 27 to November 9, 1978, to study air pollutant transport in the lower atmosphere. The tracer gas used in these experiments was sulfur hexafluoride (SF6), which was released from a tower located at a height of 115 meters (Hs), with a roughness length of 0.6 meters (z0). Additional details of the parameters used in these experiments can be found in the publications of Gryning and Lyck [19] and Gryning et al. [20].

The Prairie Grass experiments, carried out near O’Neil, Nebraska, between March and August 1956, were a groundbreaking series of studies on atmospheric pollution behavior and transport. These experiments measured the dispersion of sulfur dioxide (SO2) released from a 1.5 meter tower at five downwind distances from the source: 50, 100, 200, 400, and 800 meters. The roughness length of the area was 0.006 meters, which is critical in determining the spread of air pollutants. The experiments were extensively documented by Barad et al. [21] and Nieuwstadt et al. [22], providing essential initial insights into atmospheric dispersion processes.

Overall, the Copenhagen and Prairie Grass experiments provided essential knowledge on the transport and dispersion of air pollutants in the lower atmosphere. The Copenhagen experiments used an inert tracer gas to examine general dispersion patterns, while the Prairie Grass experiments focused on the downwind transport of SO2, providing dispersion data across a range of distances. Both experiments relied on the roughness length of the terrain, a crucial parameter in determining how pollutants spread near the ground. The findings of these landmark experiments made significant contributions to scientists’ understanding of atmospheric pollution behavior, particularly in the PBL near the Earth’s surface.

It should be noted that the computations were carried out using two distinct programming environments, namely MATLAB and Wolfram Mathematica. Specifically, the function fzero, which utilizes a combination of the bisection, secant, and inverse quadratic interpolation methods, was defined and executed in the MATLAB environment. Conversely, the FindRoot function, which combines Brent’s and Newton’s methods along with a method that approximates the Jacobian, was defined and executed in the Wolfram Mathematica environment.

To implement the analytical solution in a practical sense, we have employed a discretization technique for the atmospheric boundary layer. Specifically, we have discretized the layer into two and four sublayers, respectively. In the case of two sub-layers, the boundary layer height is discretized using the following approach: For the two sub-layers case, the boundary layer height is divided into two sub-layers, and the discretization is taken as follows: dz1=713Hmixz0 and dz2=613Hmixz0. On the other hand, for the four sub-layers case, the boundary layer height is divided into four sub-layers, and the discretization is taken as follows: dz1=315Hmixz0, dz2=315Hmixz0, dz3=715Hmixz0, and dz4=215Hmixz0. These formulas are used to accurately model the behavior of the boundary layer height in numerical simulations.

In this work, the results of the numerical convergence for concentration predictions of atmospheric dispersion models are indicated as a function of the number of eigenvalues used in the model for different downwind distances from the source. Reference is made to experiments conducted in Copenhagen (experiments 1 to 3) and Prairie Grass (experiments 1, 5, and 7).

For the Copenhagen experiments, the models used four sub-layers in the vertical and tested two different formulations for calculating the vertical eddy diffusivity, a parameter that represents turbulence and mixing in the vertical direction. Figure 1 shows the numerical convergence of normalized crosswind-integrated concentration as a function of the number of eigenvalues used in the model. The convergence is shown for four different downwind distances (200, 500, 1000, and 1500 m) from the source. The sentence indicates that for Prairie Grass, only one formulation of vertical diffusivity was used because the mixing height (Hmix) was much greater than the surface layer height (Hs), so the two formulations gave the same result.

Figure 1.

Copenhagen experiments number 1, 2, and 3: Numerical convergence of the proposed solution for two vertical turbulence formulations as a function of the number of eigenvalues at four distances using four sub-layers.

Figure 2 shows the numerical convergence of ground-level crosswind-integrated concentration for Prairie Grass experiments as a function of the number of eigenvalues at four downwind distances (50, 100, 200, and 800 m).

Figure 2.

Prairie grass experiment numbers 1, 5, and 7: Numerical convergence of the proposed solution as a function of the number of eigenvalues at four distances using four sub-layers.

Numerical simulations were conducted to investigate the influence of the vertical eddy diffusivity formulation on the normalized crosswind-integrated concentration of air pollution near a point source. Specifically, the simulations were performed for Copenhagen experiments numbered 1, 2, and 3, and the results are presented in Figure 3.

Figure 3.

Normalized concentration Eq. (24) as a function of distance using four sub-layers for Copenhagen experiment numbers 1, 2, and 3.

The figure depicts the normalized crosswind-integrated concentration cHyxz0Q at ground level as a function of downwind distance, using two different vertical turbulence formulations. The results indicate that the choice of vertical eddy diffusivity formulation has a significant impact on the concentration near the point source. Notably, the figure shows that the concentration is higher when using the first formulation of the vertical eddy diffusivity, compared to the second one. Furthermore, the figure illustrates the expected decrease in crosswind-integrated concentration with increasing downwind distance due to the dispersion of air pollutants caused by turbulent mixing in the atmosphere. In conclusion, these simulations provide valuable insights into air pollution dispersion behavior and can inform future experiments and modeling efforts.

The quality and performance of models are typically evaluated by comparing their predictions to observed (or measured) values. When observations are available, a common approach to evaluating model performance is to plot the predicted values against the observed values in a scatter diagram.

In this study, the performance of the models for the Copenhagen and Prairie Grass experiments was evaluated using scatter diagrams, which are shown in Figures 4 and 5, respectively. Figure 4 compares the predicted and observed crosswind-integrated concentrations for two different formulations of the vertical diffusivity parameter and two different numbers of sub-layers used to represent the atmospheric boundary layer (with H=2 and H=4), while Figure 5 compares the predicted and observed ground-level crosswind integrated concentrations for two different numbers of sub-layers in the atmospheric boundary layer (again with H=2 and H=4).

Figure 4.

Comparison of observed and predicted crosswind-integrated concentrations in Copenhagen (top panel) and prairie grass (bottom panel) experiments using scatter plots. The one-to-one line (y=x) and factor-of-two lines (y=0.5×x and y=2×x) are shown.

Figure 5.

Isolines of non-dimensional predicted crosswind-integrated concentration as a function of non-dimensional distance wxumeanHmix and non-dimensional depth zHmix are presented for Copenhagen experiment number 2 (top panel) and prairie grass experiment number 7 (bottom panel) at different source heights: (a) HsHmix=0.1, (b) HsHmix=0.25, (c) HsHmix=0.5, and (d) HsHmix=0.75, using four sub-layers.

The scatter diagrams show good agreement between the predicted and observed values, indicating that the models are well-parameterized. However, a visual inspection of the figures also reveals that the observed concentrations tended to be slightly higher than the predicted concentrations.

While scatter diagrams provide a useful visual evaluation, a more rigorous evaluation of model performance is achieved through statistical metrics. In this study, the statistical indices used to evaluate the models’ performance are defined in the reference [23]. These statistical metrics provide a more detailed and quantitative assessment of the models’ performance, which can be used to identify any areas of weakness or sources of error in the models. We defined the statistical indices used in our study as:

  • Normalized Mean Square Error (NMSE): A measure of the accuracy of a model or prediction, calculated as the mean square error divided by the variance of the observed data:

    NMSE=cocp2¯co¯×cp¯,

  • Mean Relative Square Error (MRSE): A measure of the accuracy of a model or prediction, calculated as the mean square error divided by the mean of the observed data squared:

    MRSE=4cocp/co+cp¯2¯,

  • Correlation Coefficient (COR): A measure of the linear relationship between two variables, usually the predicted and observed values of a model or prediction. It ranges from −1 to 1, with 1 indicating a perfect positive correlation and − 1 indicating a perfect negative correlation:

    COR=cpcp¯coco¯¯σoσp,

  • Fractional Bias (FB): A measure of the systematic bias in a model or prediction, calculated as the difference between the mean predicted and observed values divided by the mean observed value:

    FB=2co¯cp¯co¯+cp¯,

  • Fractional Standard Deviation (FS): A measure of the variability in a model or prediction, calculated as the standard deviation of the predicted values divided by the mean observed value:

    FS=2σoσpσo+σp,

  • Mean Geometric Ratio (MG): A measure of the logarithmic bias in a model or prediction, calculated as the exponential of the mean of the differences between the logarithms of the predicted and observed values:

    MG=explogco¯logcp¯,

  • Variance of Geometric Ratio (VG): A measure of the variability in the logarithmic bias of a model or prediction, calculated as the variance of the differences between the logarithms of the predicted and observed values:

    VG=explncolncp2¯,

  • Factor of Two (FAC2): A measure of the accuracy of a numerical weather prediction model, defined as the distance between the observed and predicted position of a weather system divided by the radius of the weather system. If the distance is less than or equal to twice the radius, the forecast is considered accurate and has a FAC2 score of 1. If the distance is greater than twice the radius, the forecast is considered inaccurate and has a FAC2 score of less than 1:

0.5cp/co2.

The satisfactory numerical results presented in Figure 4 are supported by the values obtained from the statistical performance measures. Upon examining these values, it is observed that they mostly fall within the range of acceptable performance for the two proposed formulations for vertical turbulent diffusivity. However, it should be noted that the simulated concentrations for experiments conducted with Prairie Grass tend to slightly overestimate the measured quantities. Specifically, when considering the turbulent diffusivity that depends on both x and z, the model produces results that are relatively more promising than those obtained with other methods.

Figure 5 depicts the spatial distribution of pollutants for two different experiments—Copenhagen experiment number 2 and Prairie Grass experiment number 7, using contour lines that represent the adimensional transverse integrated concentration predicted based on the adimensional distance for experiment 2 and adimensional depth for experiment 7, where the adimensional distance for experiment 2 is defined as wx/umeanHmix.

Advertisement

6. Conclusions

A solution has been developed for the 3-dimensional stationary state atmospheric diffusion equation, which incorporates more realistic formulations of wind velocity profiles and two formulations of vertical turbulent diffusivity. This solution was achieved through a combination of an appropriate auxiliary eigenvalue problem with mathematical induction, while a transcendental equation was developed to determine the eigenvalues for any number of subdomains. The advection-diffusion equation was solved by dividing the planetary boundary layer into discrete layers and assuming average values of turbulent diffusivity and wind velocity in each subdomain. The solution was evaluated using the Copenhagen and Prairie Grass experiments for two and four layers, and the numerical convergence of the solution was verified based on the number of eigenvalues. The results indicated good agreement between predicted and observed values, and most of the calculated statistical indices were within an acceptable range for model performance. This current model could be a promising approach for accurately predicting atmospheric pollutant dispersion and may also be applicable to other continuous flows (Table 1).

Copenhagen experiments
ModelsNMSEMRSECORFBFSMGVGFAC2
Kz=ξxφzzH=20.05330.05330.8657−0.01350.00180.96601.08851.096
Kz=φzzH=20.06600.06600.84980.04320.038971.02351.10071.055
Kz=ξxφzzH=40.07680.07680.83080.05430.16861.02731.10220.9235
Kz=φzzH=40.08920.08900.82740.09730.20651.07261.11240.8882
[24] (H=2)0.080.86−0.020.051.0
[24] (H=4)0.10.820.10.040.92
[25]0.069−0.0090.0510.9961.0551.009
Prairie Grass experiments
Kz=ξxφzzH=20.02100.02090.9809−0.0503−0.01080.90181.06450.9807
Kz=ξxφzzH=40.06130.06080.9763−0.1809– 0.05140.86161.32280.9373
[26]0.250.920.030.200.68
[27]0.040.96−0.090.130.79

Table 1.

Statistical measures for the Copenhagen and Prairie grass experiments.

References

  1. 1. Moreira D, Vilhena M, Tirabassi T, Buske D, Cotta R. Near-source atmospheric pollutant dispersion using the new GILTT method. Atmospheric Environment. 2005;39:6289-6294 Available from: http://linkinghub.elsevier.com/retrieve/pii/S1352231005006023
  2. 2. Buske D, Vilhena M, Moreira D, Tirabassi T. Simulation of pollutant dispersion for low wind conditions in stable and convective planetary boundary layer. Atmospheric Environment. 2007;41:5496-5501 Available from: http://linkinghub.elsevier.com/retrieve/pii/S1352231007003998
  3. 3. Tirabassi, T., Buske, D., Moreira, D. and Vilhena, M. A two-dimensional solution of the advection–diffusion equation with dry deposition to the ground. Journal of Applied Meteorology and Climatology. 47, 2096-2104 (2008), ISBN: 1558-8432
  4. 4. Vilhena M, Buske D, Degrazia G, Quadros R. An analytical model with temporal variable eddy diffusivity applied to contaminant dispersion in the atmospheric boundary layer. Physica A: Statistical Mechanics and its Applications. 2012;391:2576-2584 ISBN: 0378-4371
  5. 5. Kumar P, Sharan M. An analytical model for dispersion of pollutants from a continuous source in the atmospheric boundary layer. Proceedings of the Royal Society A. 2010;466:383-406 Available from: http://royalsocietypublishing.org/doi/10.1098/rspa.2009.0394
  6. 6. Vilhena M, Costa C, Moreira D, Tirabassi T. A semi-analytical solution for the three-dimensional advection–diffusion equation considering non-local turbulence closure. Atmospheric Research. 2008;90:63-69 Available from: http://linkinghub.elsevier.com/retrieve/pii/S0169809508000914
  7. 7. Pérez Guerrero J, Pimentel L, Oliveira-Júnior J, Heilbron Filho P, Ulke A. A unified analytical solution of the steady-state atmospheric diffusion equation. Atmospheric Environment. 2012;55:201-212 Available from: http://linkinghub.elsevier.com/retrieve/pii/S1352231012002567
  8. 8. Kumar P, Sharan M. An analytical dispersion model for sources in the atmospheric surface layer with dry deposition to the ground surface. Aerosol and Air Quality Research. 2016;16:1284-1293 Available comment: http://aaqr.org/articles/aaqr-15-09-oa-0549
  9. 9. Deaves DM, Harris RI. Construction Industry Research and Information Association. A Mathematical Model of the Structure of Strong Winds. London: CIRIA; 1978. 49 p
  10. 10. Mooney CJ, Wilson JD. Disagreements between gradient-diffusion and Lagrangian stochastic dispersion models, even for sources near the ground. Boundary-layer meteorology 1993;64:291-296
  11. 11. Degrazia GA, Campos Velho HF, Carvalho JC. Nonlocal exchange coefficients for the convective boundary layer derived from spectral properties. Contributions to Atmospheric Physics. 1997;70
  12. 12. Huang C. A theory of dispersion in turbulent shear flow. Atmospheric Environment. 1967;13:453-463 (1979,1). Available from: http://linkinghub.elsevier.com/retrieve/pii/0004698179901392
  13. 13. Brown M, Pal Arya S, Snyder WP. Descriptors derived from a non-Gaussian concentration model. Atmospheric Environment. 1997;31:183-189 Available from: http://linkinghub.elsevier.com/retrieve/pii/1352231096004876
  14. 14. De Visscher A. Air Dispersion Modeling: Foundations and Applications. Somerset New Jersey: Wiley; 2014. http://www.books24x7.com/marc.asp?bookid=62593
  15. 15. Pasquill F, Smith FB. Atmospheric diffusion: Study of the dispersion of windborne material from industrial and other sources. New York, NY, USA: John Wiley & Sons; 1983
  16. 16. Mikhailov M, Ozisik M. Unified Analysis and Solutions of Heat and Mass Diffusion. New York: John Wiley and Sons Inc.; 1984 Available from: http://www.osti.gov/biblio/5949945
  17. 17. Hanna SR, Briggs GA, Hosker JRP. Handbook on Atmospheric Diffusion: Prepared for the Office of Health and Environmental Research Office of Energy Research U.S. Department of Energy. Oak Ridge TN: Technical Information Center U.S. Dept. of Energy; 1982
  18. 18. Obukhov A. Turbulence in an atmosphere with a non-uniform temperature. Boundary-Layer Meteorology. 1971;2:7-29 Available from: http://link.springer.com/10.1007/BF00718085
  19. 19. Gryning S, Lyck E. Atmospheric dispersion from elevated sources in an urban area: Comparison between tracer experiments and model calculations. Journal of Applied Meteorology. 1984;23:651-660 Available from: http://journals.ametsoc.org/doi/10.1175/1520-0450(1984)023%3C0651:ADFESI%3E2.0.CO;2
  20. 20. Gryning S, Holtslag A, Irwin J, Sivertsen B. Applied dispersion modelling based on meteorological scaling parameters. Atmospheric Environment (1967). 1987;21(1):79-89 Available from: http://linkinghub.elsevier.com/retrieve/pii/0004698187902733
  21. 21. Barad M. Project Prairie Grass: A Field Program in Diffusion, Vols. I and II of Geophysical Research Paper, 59, AFCRL-TR-58-235 (ASTIA Document No. AF-152572). Bedford: Air Force Cambridge Research Laboratories; 1958 Available from: http://www.harmo.org/jsirwin/
  22. 22. Nieuwstadt F. An analytic solution of the time-dependent, one-dimensional diffusion equation in the atmospheric boundary layer. Atmospheric Environment (1967). 1980;14(1):1361-1364 Available from: http://linkinghub.elsevier.com/retrieve/pii/0004698180901547
  23. 23. Chang J, Hanna S. Air quality model performance evaluation. Meteorology and Atmospheric Physics. 2004;87 Available from: http://link.springer.com/10.1007/s00703-003-0070-7
  24. 24. Marie E, Hubert B, Patrice E, Zarma A, Pierre O. A three-dimensional analytical solution for the study of air pollutant dispersion in a finite layer. Boundary-Layer Meteorology. 2015;155:289-300 Available from: http://link.springer.com/10.1007/s10546-014-9997-0
  25. 25. Laaouaoucha D, Farhane M, Essaouini M, Souhar O. Analytical model for the two-dimensional advection-diffusion equation with the logarithmic wind profile in unstable conditions. International Journal of Environmental Science and Technology. 2022;19(7):6825-6832. DOI: 10.1007/s13762-021-03554-1
  26. 26. Moreira D, Tirabassi T, Vilhena M, Goulart A. A multi-layer model for pollutant dispersion with dry deposition to the ground. Atmospheric Environment. 2010;44:1859-1865. Available from: http://linkinghub.elsevier.com/retrieve/pii/S1352231010001524
  27. 27. Pérez Guerrero J, Pimentel L, Skaggs T. Analytical solution for the advection–dispersion transport equation in layered media. International Journal of Heat and Mass Transfer. 2013;56:274-282 Available from: http://linkinghub.elsevier.com/retrieve/pii/S0017931012007053

Written By

Mehdi Farhane and Otmane Souhar

Reviewed: 19 June 2023 Published: 18 July 2023